Handbook of yarn production
Handbook of yarn production
Technology, science and economics
Peter R. Lord
CRC Press
Boca Raton Boston New York Washington, DC
WOODHEAD
PUBLISHING LIMITED
Cambridge England
Published by Woodhead Publishing Limited in association with The Textile Institute
Woodhead Publishing Ltd
Abington Hall, Abington
Cambridge CB1 6AH, England
www.woodhead-publishing.com
Published in North America by CRC Press LLC
2000 Corporate Blvd, NW
Boca Raton FL 33431, USA
First published 2003, Woodhead Publishing Ltd and CRC Press LLC
© 2003, Woodhead Publishing Ltd
The author has asserted his moral rights.
Originally published in 1979 by the author under the title The economics, science and
technology of yarn production, this is a new, completely revised version of the book.
This book contains information obtained from authentic and highly regarded sources.
Reprinted material is quoted with permission, and sources are indicated. Reasonable
efforts have been made to publish reliable data and information, but the author and
the publishers cannot assume responsibility for the validity of all materials. Neither
the author nor the publishers, nor anyone else associated with this publication, shall be
liable for any loss, damage or liability directly or indirectly caused or alleged to be
caused by this book.
Neither this book nor any part may be reproduced or transmitted in any form or by
any means, electronic or mechanical, including photocopying, microfilming and
recording, or by any information storage or retrieval system, without permission in
writing from the publishers.
The consent of Woodhead Publishing and CRC Press does not extend to copying
for general distribution, for promotion, for creating new works, or for resale. Specific
permission must be obtained in writing from Woodhead Publishing or CRC Press for
such copying.
Trademark notice: Product or corporate names may be trademarks or registered
trademarks, and are used only for identification and explanation, without intent to
infringe.
British Library Cataloguing in Publication Data
A catalogue record for this book is available from the British Library.
Library of Congress Cataloging in Publication Data
A catalog record for this book is available from the Library of Congress.
Woodhead Publishing ISBN 1 85573 696 9
CRC Press ISBN 0-8493-1781-9
CRC Press order number: WP1781
Typeset by Replika Press Pvt Ltd, India
Printed by TJ International, Cornwall, England
Contents
Acknowledgments ....................................................................................................... ix
1
Review of yarn production ............................................................................ 1
1.1
1.2
1.3
1.4
2
Textile products and fiber production ...................................................... 18
2.1
2.2
2.3
3
Textile materials (fabrics, fibers, and filaments) ............................... 18
Natural fibers (types and production) ................................................ 22
Man-made fibers (polymer extrusion and yarn production) ............. 38
References ........................................................................................... 54
Common principles .................................................................................... 56
3.1
3.2
3.3
3.4
3.5
3.6
3.7
3.8
3.9
3.10
3.11
3.12
4
Historical basis ...................................................................................... 1
Present day conditions .......................................................................... 8
Future of the means of textile production ............................................ 9
Modern production systems ................................................................ 10
References ........................................................................................... 17
Introduction ......................................................................................... 56
Twist in strands ................................................................................... 56
Twist insertion ..................................................................................... 61
Confined and non-confined systems .................................................. 67
Twist evenness ..................................................................................... 68
Tension control .................................................................................... 69
Drawing ............................................................................................... 70
Consequences of roller errors on the textile product ........................ 76
Control of irregular flow in drawing or drafting ............................... 77
Doubling .............................................................................................. 83
Effects of shear .................................................................................... 84
Integration of sub-processes ............................................................... 86
References ........................................................................................... 87
Filament yarn production .......................................................................... 88
4.1
4.2
Introduction ......................................................................................... 88
Texturing filament yarns ..................................................................... 89
vi
Contents
4.3
4.4
4.5
4.6
4.7
4.8
4.9
4.10
4.11
5
Carding and prior processes for short-staple fibers ............................. 116
5.1
5.2
5.3
5.4
5.5
5.6
5.7
5.8
5.9
5.10
5.11
5.12
6
Introduction ....................................................................................... 155
Drawframe ......................................................................................... 155
Combing ............................................................................................ 159
Creel blending ................................................................................... 164
An industrial case study .................................................................... 165
References ......................................................................................... 167
Short-staple spinning ............................................................................... 168
7.1
7.2
8
Introduction ....................................................................................... 116
Opening line ...................................................................................... 118
Bale preparation ................................................................................ 119
The first stage of blending and opening .......................................... 121
The process of disintegration of fiber clumps ................................. 122
Condensation ..................................................................................... 123
The process of cleaning .................................................................... 125
Intimate blending .............................................................................. 129
Fiber flow .......................................................................................... 133
Carding .............................................................................................. 136
Waste control ..................................................................................... 149
Safety ................................................................................................. 153
References ......................................................................................... 154
Sliver preparation ..................................................................................... 155
6.1
6.2
6.3
6.4
6.5
7
Real twist texturing ............................................................................. 90
False twist texturing ............................................................................ 92
Draw-texturing ................................................................................... 102
Stuffer box texturing ......................................................................... 104
Air-jet texturing ................................................................................. 106
Other texturing techniques ................................................................ 110
Industrial filaments ........................................................................... 113
Silk filaments and staple yarns ........................................................ 113
Morphology and dyeing .................................................................... 114
References ......................................................................................... 114
Ring spinning .................................................................................... 168
Open-end spinning ............................................................................ 185
References ......................................................................................... 203
Long-staple spinning ................................................................................ 205
8.1
8.2
8.3
8.4
8.5
Introduction: Effects of lengthening the staple ............................... 205
Wool fibers and their preparation .................................................... 206
Worsted systems ................................................................................ 213
The woolen system ............................................................................ 220
Bast fiber spinning processes ........................................................... 231
References ......................................................................................... 232
Contents
9
Post-spinning processes ............................................................................ 234
9.1
9.2
9.3
9.4
9.5
9.6
10
Yarns of complex structure ............................................................... 260
Processes using modified twist ........................................................ 261
Compact spinning .............................................................................. 261
Air-jet spinning ................................................................................. 263
Sirospun yarns and process .............................................................. 268
Hollow spindle spinning ................................................................... 270
Self-twist spinning ............................................................................ 271
Twisted self-twist yarns and processes ............................................ 274
References ......................................................................................... 275
Quality and quality control ..................................................................... 276
11.1
11.2
11.3
11.4
12
Winding ............................................................................................. 234
Yarn joining ....................................................................................... 245
Ply yarns ............................................................................................ 250
Automation ........................................................................................ 253
Two-for-one twisting ......................................................................... 255
Customer concerns ............................................................................ 257
References ......................................................................................... 259
Staple systems and modified yarn structures ........................................ 260
10.1
10.2
10.3
10.4
10.5
10.6
10.7
10.8
11
vii
Quality ............................................................................................... 276
Quality control .................................................................................. 278
Yarn evenness .................................................................................... 291
End-breaks and quality ..................................................................... 298
References ......................................................................................... 300
Economics of staple yarn production ..................................................... 301
12.1
12.2
12.3
12.4
12.5
12.6
Yarn economics ................................................................................. 301
Productivity ....................................................................................... 303
Quality and economics ...................................................................... 306
Cost minimization ............................................................................. 308
Operational factors ............................................................................ 313
International competition .................................................................. 315
References ......................................................................................... 316
Appendices ................................................................................................ 317
1.
Calculations I: Elementary theory ....................................................... 317
2.
Calculations II: Worked examples ....................................................... 329
3.
Advanced topics I: Air conditioning and utilities ............................... 341
4.
Advanced topics II: Testing of textile materials ................................. 350
5.
Advanced topics III: Staple yarn structures ........................................ 373
6.
Advanced topics IV: Textured yarn structures .................................... 383
7.
Advanced topics V: Blending of staple fibers..................................... 389
8.
Advanced topics VI: Drafting and doubling ....................................... 407
9.
Advanced topics VII: Yarn balloon mechanics ................................... 427
10. Advanced topics VIII: Topics in rotor spinning .................................. 453
Index ........................................................................................................................ 465
Acknowledgments
Grateful acknowledgments are made to the many friends and colleagues who have
read various parts of this script and given very helpful and constructive criticism.
These include Charles Chewning, David Clapp, Philip Dabbs, Yehia E1 Moghazi,
Wally Johnson, W Oxenham, Jon Rust and W C Stuckey.
Acknowledgments are also made to UMIST, NC State University, Cotton
Incorporated, the many different commercial organizations in a variety of fields with
whom I have been associated and the many people involved. It has been a great
experience and pleasure. Much of what I have to say in this book has its origins in the
many discussions and shared fieldwork experiences. This embraces too large a number
of people to name individually but I take the opportunity to express my thanks to all
of you. I hope that by sharing these experiences, I will repay the help that I have
received during my career in this fascinating field.
I would be remiss in not acknowledging the wonderful forbearance of my wife
Mavis during these last years. Only someone that cared so much could have tolerated
the late arrivals for meals, the relegation of important social responsibilities and the
overwhelming obsession with the project. I dedicate the work to her in celebration of
our 56th wedding anniversary.
1
Review of yarn production
1.1
Historical basis
1.1.1 Historical background [1]
The long reach of history shows how prosperity varies as civilizations have waxed
and waned. The course of prosperity has been bumpy and there are dangers in
extrapolating the future based on the short-term past. Successive centuries have seen
fundamental changes of varying types. Greenwood [2] outlines steps related to yarns
and textiles in the first two millennia and points out the extraordinary fineness of the
materials that have been made. He also discusses some of the developments that have
improved the productivity of the manufacturing systems and reduced the costs over
the centuries. The eighteenth century saw a financial revolution, the nineteenth saw
the industrial revolution, and the twentieth saw the information revolution.
The history of humanity contains many references to textile materials because
they were, and still are, part of the fabric of our lives. Consequently, the history of
fibers is one of the traceable threads in the story of yarn production. A second thread
concerns the extraordinary developments of the industrial revolution. There were
gigantic steps in productivity of both people and machines. Another thread concerns
the developing economic environment that has surrounded these changes. Thus, let
us first make a brief survey of the history of some important fibers.
1.1.2 A brief history of silk
The origin of silk is found only in legend and fable; certainly it was used in the time
of Emperor Huang Ti in China in the third millennium BC. Sanskrit literature refers
to silk in India in the second century BC and the Old Testament also refers to it. When
it percolated to the West, it was as valuable as gold on a weight-to-weight basis.
Roman Emperor Justinian tried to monopolize the trade (unsuccessfully), smuggled
silk worms to Constantinople (c. AD 550) and started sericulture there. Byzantine
silks became world famous. The Moors established sericulture in Spain and so the
production of silk spread. It reached northern Europe in the fifteenth century and the
2
Handbook of yarn production
western hemisphere in the sixteenth, although it failed at first. However, the strong
luster and ability to take brilliant dyes made silk very attractive. The peak of activity
was after World War I and by 1919 the price had risen to US$21/lb; that is equivalent
to over $1200/lb in the currency of 2003. Once fine man-made fibers entered the
market, the price and the demand for silk dropped; but there is still an important
market in some areas of the world.
Perhaps the early inventors of synthetic fibers were influenced by the knowledge
of the manner in which silkworms, spiders, and other creatures extruded filaments.
Doubtless, they were also impressed by the extraordinary properties of these naturally
extruded fibers. Such inspiration was probably very important in determining the
future of fiber production.
1.1.3 A brief history of bast fibers
Bast fibers are derived from the stems of various plants.
Cultivated flax [3] probably originated in the Mediterranean region; certainly it
was used in prehistoric times. It was found in Stone Age dwellings in Switzerland, the
ancient Egyptians used it, and references to it are sprinkled throughout historical
writings. It has been used both for its fiber and for its seed. The fiber is used to make
linen cloth, and the crushed seed yields linseed oil, long used for the preservation of
leather and wood. Until the eighteenth century, linen manufacture was widespread in
the domestic industry of European countries. The development of cotton processing
and the great inventions of the industrial revolution dealt an almost fatal blow to this
erstwhile prevailing industry.
Jute fiber was largely unknown in the West until the eighteenth century, but it was
in common use in Bengal before then. There was resistance to its use because blending
it with hemp or flax was regarded as adulteration. In the nineteenth century, the
Dutch government replaced linen coffee bags with jute and this gave an impetus to
use it in the West. Research was carried out in Dundee, Scotland, which became a
recognized center of yarn production. Also, much of the production was in what are
now Pakistan and India. (Strangely, after partition in 1947, India had the jute processing
resources, and the bulk of the corresponding agricultural producing sector was in
Pakistan: Jute played a prominent role in the development of trade relations between
the two countries.) It was attractive because it was strong, bulky, and cheap. However,
in more recent times, the increasing use of polypropylene for cotton-bale wrapping,
carpet backing, sacking, and other products has decimated the jute industry.
Hemp fiber is thought to have originated north of the Himalayas and was well
known in China in the second millennium BC. It was brought to the Americas in the
sixteenth century and, by the twentieth century, was being grown throughout the
world. The plant not only produces fibers but also narcotics. Some species of hemp
produce little in the way of narcotics, but many countries make the growing of it
illegal for social reasons.
1.1.4 A brief history of wool fibers
The use of wool for clothing dates back to antiquity. Outstanding properties of
wrinkle resistance, moisture absorption, warmth, and tendency to felt, have given it
a role, not only in apparel, but also in blankets, upholstery, and floor coverings.
Babylonia is translated by some as meaning ‘the land of wool’. It is known that the
Review of yarn production
3
Phoenicians traded wool fabric during the first millennium BC [3]. The Ancient Romans
established wool factories to supply their army; the fame of these factories was
spread by the travels of Roman soldiers. In Britain, the wool flocks were scattered by
the incoming Saxons and the wool trade there then went into decline. The Normans
re-established the trade and it developed for a time, although there seems to have
been little progress through the dark ages; it was not until after the seventeenth
century that structural changes started to occur. After many struggles over restraints
in trade, wool was very important in England in the eighteenth century. Spain too was
a major producer but its government had enforced rigid restrictions on the export of
fleeces at about that time [4].
In times of rapid technological change, many are left behind. Mechanization in the
Low Countries and Britain in the nineteenth century permitted spinners in these
regions to out-produce those who had not embraced the emerging technologies. There
was then a vast opening-up of the supply of raw wool from the western and southern
hemispheres. It was the combination of a plentiful supply of raw material and high
productivities of people and machines that produced the displacement of the centers
of production and sites of the markets changed also. Now, many of the industrial
companies then formed have, in turn, been overtaken by new technology and economic
changes. The development of synthetic fibers and new processes has created a new
situation; the market in wool has declined somewhat even in the last decade. Despite
this, the world consumes about 1.5 million tons of wool per year, and its value is
greater than the weight might suggest. Australia abolished its price support in the
1980s and prices globally were determined more than before by market forces. In the
decade centered on 1990, prices plummeted [5], but supply is now in better balance
with demand and there is hope for expansion. China is now a large consumer.
A remarkable feature of wool is its ability to recover from deformation over a time,
and this gives apparel made from the fibers attractive crease-shedding properties.
Also, the rate at which the fiber takes up and disperses moisture is such that it gives
clothes made from wool good comfort properties. These inherent properties give
wool an attraction that is likely to guarantee it a place in the world market; the main
question is how much of that market it will retain.
1.1.5 A brief history of cotton fibers
The use of cotton fibers has been traced back to as far as 3000 BC. Yarns were found
in the ruins of Mohenjo-Daro, a city in the Indus valley [2]. Cotton has been known,
cultivated, and worked in India since the earliest historical periods. A Hindu Rigveda
hymn (c 1500 BC) mentions cotton, and Herodotus (c 450 BC) is said to have mentioned
‘wild trees bearing fleeces as their fruit’. Ancient Egyptians were known to have
grown and spun yarns in the seventh century AD. When the Spaniards arrived in
America they found cotton being used to make cloth. Cotton was found in prehistoric
pueblo ruins in Arizona, and cotton grave cloths from pre-Inca Peru are still in
existence.
Cotton has remarkable durability in the marketplace; it filled a major role in the
industrial revolution and it has formed an alliance with man-made fibers in more
recent times. Therefore, it is perhaps best if further discussion of the history of cotton
is left to unfold with some of those events.
4
Handbook of yarn production
1.1.6 A brief history of man-made fibers
Ideas about synthetic fiber processes were expressed by Robert Hook (1775) and
René de Reaumur (1734); Louis Schwabe extruded glass fibers in 1842. Much of the
early work was to develop a means of ‘liquefying’ cellulose to permit extrusion. In
1846, C F Schoenbein prepared nitrocellulose, and George Audemars patented a
process for making a material related to rayon from nitrated wood in 1855. Sir Joseph
Swan coagulated nitrocellulose solutions to produce fibers but he was interested
mostly in producing filaments for electric light bulbs. The stage was therefore set for
Count Hilare de Chardonnet to begin commercial production, in 1891, of filaments
coagulated in heated air, from a nitrocellulose solution derived from mulberry leaves.
Louis Henri Despeissis then developed a cuprammonium solvent for cellulose itself,
from which filaments could be coagulated in sulfuric acid. Another important milestone
was when C F Cross and E J Bevan patented (1892) a viscose rayon solution resulting
from dissolving cellulose xanthate in dilute sodium hydroxide. In 1902, Max Mueller
discovered a way to convert cellulose xanthate into regenerated cellulose and the
production of viscose rayon yarn could then start at Marcus Hook in Pennsylvania,
USA, in 1911. (It might be added that the pressures from environmental concerns at
the end of the twentieth century have led to the closure of some plants making
products of this nature.)
Synthetic polymers (which are large molecules, or ‘macromolecules’) were developed
as a result of research into the properties of large molecules starting in 1926. This is
an early example of the commercial exploitation of organized scientific research.
Wallace H. Carothers and his associates found that they could draw out filaments
from long-chain polymers. Such filaments were extruded and toothbrush bristles
were made from them. Soon after that, polyamide filaments for knitted hosiery
entered limited production, and in 1940 wider production began of the first truly
commercial filament – nylon. Nylon is a polyamide and the idea of extruded
macromolecules has since expanded to include many other chemical types. Polyester
was developed in the 1950s and Brunnschweiler and Hearle have collated an interesting
account of this development [6]. Even after the chemical and mechanical problems
(with excessive discontinuities in extrusion) had been reasonably well worked out,
there were still difficulties with developing appropriate fiber crimping and finishes.
There were also difficulties with static electrification, dyeability, oily soiling, pilling,
dye sublimation during ironing, hole melting, and other problems in fabrics. But, by
1952, production had begun and there was a boom that lasted for five years; then the
pace of market development moderated as the producers of natural fibers formed
their own marketing organizations. The excesses of a multitude of small, single-knit
fabric producers also caused a condition of oversupply of knitted fabric. Since that
time, however, there has been a continuous expansion of the markets for all fibers. In
those markets, polyester has expanded to some 10 million tons /year, with acrylic
fiber running at about one-third of this. It is noteworthy that in staple production, the
ratio between polyester and natural fibers still holds at roughly equal proportions.
The centers of production of man-made fibers have spread throughout the world.
Man-made fibers have been produced in filament and staple forms. In the early
days, the cellulosic filaments were difficult to texture, and the shiny, slick surfaces
were not in universal demand. Many were cut up into staple and blended with natural
fibers. As synthetic filaments began to appear, means of texturing them followed
closely and a successor to the silk throwing industry, aimed at the apparel industry,
began to appear in advanced countries. Technical filaments rapidly penetrated the
Review of yarn production
5
industrial market because of the strength of the materials, and heavy textured nylon
yarns made a large penetration of the home furnishings market, especially in carpets.
Also, blends of synthetic staple and natural staple fibers began to appear. Of these,
polyester/cotton and polyester/wool blends have become significant raw materials for
yarn makers.
1.1.7 Historical development of the economic environment
Religious persecution of the Huguenots in mainland Europe before the seventeenth
century caused them to flee to England and they carried their knowledge of textile
manufacturing with them. The village of Worstead in Norfolk gave its name to the
worsted process at that time. The business skills of these and others led to the
development of the idea that credit is as good as cash. The outcome of this was the
founding of the first publicly financed companies, leading to an explosion of business
ventures that swept the inventions into the industrial revolution. Others similarly
exported their skills over succeeding decades and centuries with the result that
technological and business environments spread through the world. Writers have
referred to the industrial revolution as though it occurred in a flash. In reality, it was
spread over the eighteenth and nineteenth centuries, during which time there were
ups and downs in all the economies concerned. Similarly, the newly termed ‘information
revolution’ will evolve over a considerable timespan.
The success of the industrial revolution stems from the opening of new markets.
For example, in the initial phase, cotton fibers were very expensive in markets where
cotton did not grow naturally. Cotton, at the beginning of the industrial revolution,
cost the equivalent of roughly US$100/lb in 2003 currency. The alternatives used by
many were wool and bast fibers; the yarns were coarse and the fabrics heavy. Only
the rich could afford to pay more than the equivalent of, say, $100/lb for fine cotton
yarns. Cotton in Europe at the beginning of the twentieth century, cost roughly the
equivalent of $10/lb in modern currency. In 2003 it is less than $1/lb. Reductions in
the cost of fiber did not happen accidentally. Expanding demand, improvement in
agricultural techniques, mechanization of harvesting and ginning, and, ultimately, the
rise of man-made fibers all put downward pressure on fiber prices.
The beginnings of the industrial revolution involved extraordinary inventiveness,
availability of capital, and much human exploitation. Development of machinery did
little to make the life of the mill worker easier at the turn of the century; much of the
benefit went to the mill owners and traders. Many inventors were also excluded from
benefit. The short-term result in the nineteenth century was that the cost of the goods
produced was sharply reduced and there was a worldwide expansion of trade. In the
early stages, goods were transferred from a rich economy (Group A) to others (Group
B) and wealth in other forms flowed in the opposite direction as shown in Fig. 1.1. As
Goods
A
B
Wealth
Fig. 1.1 Historical flow of wealth
6
Handbook of yarn production
Cost (log scale)
the development continued, Group A expanded to include a range of nations and
economies so that the simplistic model given here became less applicable, but the
general idea is the same.
Long-term results of the worker exploitation in the Group A economies were the
rise of socialism and the associated rise in wage costs during the twentieth century
(Fig. 1.2). This widened the economic gap between Groups A and B; also, there was
a marked decrease in fiber prices as the market expanded. These two factors played
important parts in the economic changes throughout the world. Of course, there were
short-term fluctuations in these costs; therefore Fig. 1.2 shows only trend lines.
It is not surprising to find that a growing number of the Group B economies
desired to join the industrial nations and set up production units under their own
control. This was especially so during the second half of the twentieth century. The
element of widely differing wage levels became an important factor in the competition,
and eventually in the net flow of textile goods. To combat the differences in wages,
machinery was then developed to reduce the need for human intervention and improved
to increase machine productivity as indicated in Fig. 1.3. The increases in productivity
now seem to be leveling off and perhaps we should not expect the same massive
changes in the present century that we have experienced in the last.
The enhancement of machine productivity was not the only contributor to the
reduction in costs. At the beginning of the twentieth century, it was normal to have six
or seven stages of processing between carding and spinning in the cotton industry.
Nowadays, some mills work with as little as one stage between those processes. The
reduction in the number of stages gives at least two benefits. First, it reduces the
capital cost components. Secondly, it reduces the need for transfers of the textile
product from one machine to the next. Formerly, these transfers were manual and
10 Cotton fiber, $/lb
1.0 Wages, $ /hr
1850
Fig. 1.2
1900
1950
Year
2000
Costs in 1994 US$ in an ‘A’ economy
Productivity, log scale
(tons/spindle year)*
10
1.0
0.1
0.01
1900
1950
Year
*Estimated values
Fig. 1.3
2000
Changes in machine productivity
100
200
24/ l Cotton yarn
50
100
50
20
10
20
Group B
10
Group A
1950
Fig. 1.4
7
Year
5
(operator hrs /100 lb)
Productivity, log scale
(operator hrs /100 kg)
Review of yarn production
2000
Changes in operator productivity
thus accounted for much of the need for human resources. Modern automated transfer
systems are now available, which change the balance between the cost categories of
labor and capital. The transfers are still a significant cost component; reducing the
number of them plays a part in the economies of the modern technology. Technical
developments include automation, automatic handling, and reduction of the number
of process stages as well as enhanced productivity of equipment. The rate of adoption
has varied throughout the world and, for the present purpose, rates appropriate to the
categories mentioned earlier are indicated in Fig. 1.4 to illustrate the point. The
normalized variable quoted in the diagram in metric units (known colloquially as
‘HOK’) is a measure of the productivity of the workers in the factories.1 If we
multiply the wage rate by the HOK variable we get the labor cost per 100 kg of
product. Labor cost per unit weight has changed less than might be expected over the
last century after discounting inflation and allowing for changes in other cost
components. This is despite the steep rise in wage rate shown earlier. The cost per
unit weight has been low in the Group B economies and this has given them an
advantage in the past.
However, as the HOK shrinks, so do the effects arising from the differences in
wages; other cost factors then tend to predominate. As wealth flows to the lesserdeveloped countries, the standards of living improve and wage rates rise. As wealth
flows from Group A economies, there is a trend for a particular industry to decline,
sometimes to the detriment of the local standard of living. In other words, there is a
tendency towards equalization of wages. Of course, there are more than two categories
of economy in the world and the spectrum of wage rates is wide. Nevertheless, the
trend still exists, with some nations moving up the scale and others moving down; it
is a fluid situation with changes certain over the coming years. Losses in manufacturing
jobs due to automation also have had, and will have, a marked effect. Automation not
only affects costs of production but also affects the tastes and the ability of the
remaining workers to buy textile goods. The markets tend to be more widely distributed
and their character changes. The cost of fibers has changed, especially in the last half
of the twentieth century, and there are a number of reasons for this. There was more
attention given in the second half of the century to civilian matters than in the first.
1 HOK = operator hours per 100 kilograms of product and OHP = operator hours per 100 lb of
product. The acronym ‘HOK’ appears to be illogical because it is derived from a language other
than English.
8
Handbook of yarn production
Much of the first half was spent in war and depression, and the markets did not
realize sustained expansion. The economic expansion after 1950 probably played a
large part in reducing costs because of improved efficiencies of the organizations and
equipment. Also, man-made fibers became a larger competitor to natural fibers.
A range of man-made fibers became commercially viable over a number of markets
in the second half of the century. The sales were accompanied by technical service
and research that developed more new markets and facilitated competition in many
traditional ones. As time progressed, the production of man-made fibers spread
throughout the world and is now ubiquitous. Producers of natural fibers have combined
to form various organizations that emulate the research and service provided by the
man-made fiber companies. But if the price of a fiber falls, as it did with wool, then
the service from such organizations becomes more restricted. There have been cutbacks
over a range of such organizations. In general, as fibers become more nearly perfect
commodities, less is spent by fiber producers on research. Eventually, they are less
able to give the same technical support to the textile producer as before. It is therefore
likely that textile producers will have to look for other resources and this may have
an impact on some companies’ ability to compete.
1.2
Present day conditions
1.2.1 Costs and sales
In yarn production, labor costs are only part of the total cost. Livingston [7,8] states
that US labor costs formed 14% of the total in 1992; the largest component was said
to be that of fiber costs, which comprised approximately 50% of the total. However,
the percentages vary from place to place in the world. Energy costs in yarn production
vary with the product but, for example, can run at up to some 10% of the total. These
costs rise with speed and count. It can be expected that power requirements will rise
in Group B economies as more plants run at higher speeds and install air conditioning.
Also, amortization costs rise with investments in equipment. Again, there is a trend
for equalization of the costs between the various economies. Where the main flow of
goods is global, shipping costs become increasingly important because they impose
a premium of perhaps some 10% on the transoceanic shipper of the goods. An
additional hindrance to trade is caused by tariffs and quotas. The fairly recent international actions expressed by GATT (General Agreement on Tariffs and Trade) and
NAFTA (North American Free Trade Agreement), and others to follow, are likely to
reduce these barriers and there are hopes for an enlarged market. There is emerging
evidence that the move of parts of the textile industry from Group A to Group B
regions has been hastened by the freer market. Some erstwhile Group B regions have,
indeed, become Group A regions. Relative currency changes also affect the issue.
Sales are not determined by price alone. Quality of the product and service also
affect the issue. Most textile products go through a chain of sales transactions before
reaching their final destination. These intermediate transactions are between
professionals, and technical quality becomes important. Of course, the requirements
of the purchasing public have also to be considered. The point is that considerable
investment has to be made in appropriate testing equipment, and care in testing
becomes essential to satisfy buyers. Probably the most important aspect of service is
delivery of goods at the specified times. To get the best advantage, quality and
service have to be managed efficiently.
Review of yarn production
9
History leads to think that the most important factor for success is to recognize
expanding markets and plan accordingly. Mechanical inventions may have a relatively
small effect on the future competitive position. Other technologies, such as
telecommunications and computing, are likely to have a greater effect. Nevertheless,
the need to operate a mill in the most economic manner is still a paramount consideration,
and high productivity machinery has to be used for major installations. In addition,
there is a great need to make products of a quality that will satisfy the market; this
involves quality control, which becomes ever more sophisticated. Cost and quality of
the product have to be carefully balanced for each market to achieve a competitive
position without which the enterprise will fail.
1.2.2 World market for yarns
According to Thomas et al. [9], the market for spun yarns will be dominated by
cotton. At the time of writing, the share held by cotton is about two-thirds of the
world market and this has been stable in recent years. Nevertheless, this is not to say
that the market has remained unchanged; on the contrary, shifts in consumer demands
and preferences, cost structures, and geographic migrations of the industry are powerful
agencies for change.
Europe has suffered losses in production capabilities whereas Asian output has
soared. American output has increased but the character has changed. There is
consolidation amongst the companies that might be seen as evidence of the sorts of
pressures that have affected Europe. However, Europe is still the world leader in the
smaller market of long-staple spinning. The production of cotton is still very strong
in the USA and this is one of the reasons why the industry there has maintained
stability. Production of polymers and man-made fibers and filaments has dispersed
through the world and, again, the production in Asia has made remarkable strides and
is affecting Western markets.
1.3
Future of the means of textile production
One reason for the reduction in HOK over the years is that the productivity of the
machines has increased. Greenwood [2] quotes HOK values ranging from 12 500 in
Neolithic times, through 3120 in ‘pre-fourteenth century’, to 0.63 for open-end (OE)
spinning in the 1970s. In staple spinning, the mule was superseded by the ring frame
and then by newer technology. The move to rotor spinning and other new technology
in the USA and in some other areas has been highly significant. The productivity of
the fastest machines has escalated rapidly but it is difficult to imagine how the pace
can continue. There are signs that the productivity curves are flattening and they
seem to be approaching maximum values asymptotically.
Next let us turn to materials handling. At the beginning of the twentieth century,
the whole process consisted of a myriad of steps, with human intervention at each
one. Gradually the number of steps has been reduced and automatic handling has
become common. Automatic handling takes several forms. It ranges from the pneumatic
transfer of fiber, to the use of robots to carry packages between machines. It follows
that these developments have also contributed to the reduction in labor. Again there
are limits; as we approach the irreducible minimum number of stages and automate
the transfer of textile material, there is little to be gained in possible labor cost
10
Handbook of yarn production
reductions. However, there might well be other advantages. Thus, for this reason, the
HOK curves shown earlier cannot be extrapolated too far. Nevertheless, these
considerations imply that improvements in the technology of production will play a
diminishing role in deciding the partition of the markets.
1.4
Modern production systems
1.4.1 Some comments on the mechanics of fiber structures
Some knowledge of the properties of the fiber, and of the yarn as well, is useful to
understand the difference between various products. At this stage, suffice it to consider
only the mechanical properties of stiffness and bulk. A coarse fiber (i.e. one having
a high linear density) is stiffer than a fine one. Consequently, fabrics made from the
coarser fibers often feel harsh and prickly. Thus one can understand the drive to use
fine fibers that give a softer ‘hand’ to fabrics. However, if carried too far, the use of
ultra fine fibers can lead to difficulties with the production of nep (a fiber fault
caused by fiber knotting or tangling with itself to yield a tiny ball of fiber which may
take up dye at a different rate). Next, consider fiber bulk, which is affected by fiber
crimp or convolution. Crimp typifies the extent of zigzag, helical, or other non-linear
shape of the fiber. The greater the crimp, the more volume it takes up and the more
‘bulky’ is the yarn. Figure 1.5 is intended to show the effect of texture in regard to
bulk. Figure 1.5(a) indicates a series of parallel fibers, which could be easily compressed
to form a strand of very little bulk. Figure 1.5(b) shows similar fibers which have
been induced to curl into helical configurations, lying beside one another like a series
of bedspring coils to occupy a much greater volume than before. A bulky yarn made
up into fabric produces a material with good insulation properties. Fine, bulky fibers
produce fabrics that feel warm and soft. Fine, non-bulky fibers produce silk-like
fabrics. On the other hand, coarse fibers are often used to make carpets because the
carpet tufts stand out from the backing and can be loaded at their ends in normal use
without buckling too severely. A fine fiber would buckle and change the appearance
in the loaded areas of the carpet. Thus, the use of coarse fibers helps to reduce the
appearance of tread marks on the surface of the carpet.
(a)
(b)
Fig. 1.5
Volume occupied by fibers
Review of yarn production
11
1.4.2 Filament production
Fiber production before the eighteenth century was an agricultural undertaking with
the result that, except for silk, the fibers available were of a relatively short finite
length (we call these staple fibers). Yarn production systems of antiquity were mostly
staple yarn systems. In modern times the range of raw material has expanded and
synthetic fibers have become available. These so-called man-made fibers are supplied
in staple form and also as ‘continuous’ filament. Thus, in modern times, there is a
range of yarn making technologies which did not exist earlier. This range continues
to expand, perhaps at a declining rate, as economic factors other than machine
productivity take precedence.
Extruded filament yarn manufacture is a short, mechanical process involving only
one or two steps. Yarn is extruded and drawn to approximately the right ‘size’; it then
is often textured to give the final product. A schematic drawing of a simple melt
extrusion system is shown in Fig. 1.6. It shows only a rudimentary polymer chip feed
system. A practical system may have a complex liquid polymer feed and two or more
draw zones in a single spinline. There are a variety of alternative systems within this
broad category of filament production. Although the production of filament yarns
appears deceptively simple, there are complexities. The processing conditions have to
be very carefully monitored and controlled because heat, humidity, and mechanical
stress affect the polymer in a way that affects the dyeability of the final product. Thus,
it is imperative that those in charge understand the problems which can arise due to
the chemistry and molecular structure of the polymers from which the fibers are
made. This is in contrast to the mechanical complexities of staple processing.
Staple yarn manufacture is much more complex from a mechanical standpoint; it
involves many stages of processing before the products are ready for shipping. There
Polymer input
Melt
Extrusion
Filaments
solidify &
cool during
transit
Draw
Wind
yarn on
cone or
cheese
Fig. 1.6
Simple polymer chip feed system for the production of filament yarn (Note: practical
systems are more complex)
12
Handbook of yarn production
are a variety of staple spinning systems available, but broadly they can be categorized
as short- and long-staple systems. Short-staple spinning is the logical development of
the cotton spinning of history, but the range of fibers has increased dramatically in
this century. Long-staple spinning has a heritage of spinning wool and bast fibers; but
in recent times, the range of long-staple fibers has also increased markedly. A comparison
of various systems is given later in Table 1.1.
1.4.3 Textured yarn production
In yarn production, polymer is supplied either directly from the chemical reactor or
as polymer chip. The polymer is fed to an extruder in which a rotating screw or auger
transports the input material through the extruder barrel and pressurizes it; as the
polymer passes through the barrel it is melted or maintained in the molten condition.
The extruder changes the form of the molten polymer, from a relatively slowly
moving mass to the high speed thin jets of polymer which form the yarn. It is metered
and filtered before passing through the spinneret, which contains one tiny hole for
each filament. The emerging filaments cool rapidly and solidify; they are also ‘drawn’
by taking them up at a faster rate than that of the supply. Drawing is a very important
part of the process because it stabilizes the molecular structure and strengthens the
yarn by improving the molecular orientation.
The main idea in most texturing systems is to heat set the filaments into some sort
of crimped or convoluted form, such that each filament is held as separate from its
neighbors as possible. In this way the yarn contains the many air pockets needed to
produce insulation properties, permeability, and softness. Furthermore, the yarn now
occupies a greater volume, which is also very important since the purpose of most
textile materials is to cover some underlying strata; the greater the bulk, the better the
cover. Also the yarn becomes more extensible and this, too, is an added attraction. It
is possible to get various combinations of stretch and bulk. For filaments (such as
rayon) that cannot be heat set, it is possible to tangle the fibers to lock them mechanically.
Table 1.1
Typical process schedules
Fiber processes
Nylon
filament
Nylon
tow
Polyester
staple
Cotton
Wool
Extrusion
Drawing
Winding
Extrusion
Drawing
Stretch-break
Extrusion
Cutting
Crimping
Baling
Harvesting
Ginning
Shearing
Sorting
Scouring
Baling
Baling
Mill processes
Texturing
Winding
Drawing
Drawing
Roving
Ring spinning
Winding
Opening
Carding
Drawing
Drawing
Roving
Ring spinning
Winding
Opening
Cleaning
Carding
Drawing
Drawing
Roving
Ring spinning
Winding
Opening
Cleaning
Scouring
Carding
Drawing
Drawing
Roving
Ring spinning
Winding
Typical yarns
Typical uses
Textured
Hosiery
Staple
Carpet
Staple
Apparel
Household
Staple
Apparel
Household
Staple
Apparel
Carpet
Review of yarn production
13
An example of this is air-jet texturing. Sometimes it is desirable to combine air-jet
with false-twist texturing. Air-jet texturing gives a product that is nearer to a staple
yarn than is a false-twist textured yarn. It has much of the hand and appearance of the
staple product. False-twist machines with built-in air-jets are now becoming common.
1.4.4 Tow and man-made staple fiber
Man-made staple fibers are made from tow, which is extruded in the same basic way
as with filament yarns; however, the number of filaments involved is vastly larger.
The linear density of the filaments in the tow depends on the end use. A large
extruder supplies a number of spinnerets and several extruders are ganged together to
produce a thick rope of filaments. These filaments must be fully drawn before they
are cut into staple or shipped to the spinning mill. The major production of most tow
makers is cut within the organization and the product is sold as staple fiber. However,
some tow is sold to those mills which elect to stretch-break or cut their own staple.
Usually, such mills make long-staple yarns. Where the fiber maker cuts the material,
great care is needed to blend very large volumes of material to ensure uniformity of
the product over long periods of time. Care also is taken when the fiber makers’
processes are altered in any way because the slightest change can cause tremendous
difficulties in the mills. The fibers are batched in so-called ‘merges’ so that changes
can be equalized and controlled. This is not to say that there are never variations in
fiber properties, but rather that they are sensitive to change and that great technical
expertise is needed to control them.
For long-staple processes, where the mill chooses to convert the tow within the
mill, different standards apply. There is no longer an opportunity to merge batches in
gigantic blending operations. Each tow supplied to a mill must stand on its own merit
and the tolerances on the filaments have to be even more strict than with general tow
production because there is little opportunity for doubling; thus the cost is higher.
Another factor is that even the modern stretch-breaking machine used for converting
tow to staple is limited in the linear density of tow that it can handle. The tows
supplied to mills are often much lighter than those used internally by the fiber maker.
A simple calculation illustrates the point. If the strength2 (tenacity) is, say, 35 cN/tex
and the tow has a linear density of 106 tex, the breaking strength of the tow is 0.35
× 106 Newtons (over 30 tons). Thus, the loads needed to break the fibers in a filament
tow are very high and the stretch-breaking machine has to be very robust. Such
machines are expensive. Even the lightest tows used today might have a quarter of a
million filaments in the cross-section.
1.4.5 Staple yarn production
Staple fibers usually arrive at the mill in compacted bales containing about 500 lb of
fiber. The bales are stored according to fiber classification; this makes blend component
selection easier. Fiber supplies might come from brokers who deal in natural fibers,
from synthetic fiber makers, or from both. In the latter case the product is often
referred to as ‘blended yarn’ although, in reality, all staple yarn is blended. An
example of a flowchart for cotton and melt spun staple fibers is sketched in Fig. 1.7
2 A tex is a measure of ‘thickness’ or linear density equivalent to 1 g / km
14
Handbook of yarn production
Tow production
Grow
Harvest
Draw
Wind
Gin
Bale
Grade
Cut, oil
& crimp
Staple
fibers
Bale
Warehouse
Open
Clean
Card
Draw
Draw
Spin
Wind
Staple yarn
Fig. 1.7
Staple fiber processing
and any combination may be used according to the market being served. Somewhat
similar routes apply to wool/man-made fiber blends but in this case there has to be a
wet process stage somewhere to remove the wool grease. This is increasingly being
done in the agricultural sector.
1.4.6 Short-staple yarn production
Short-staple yarns are produced from bales of short staples, which are delivered to
the mill by the fiber suppliers. The yarn maker ‘opens’ the bales to produce a flow of
more or less discrete fibers, which are then combined into a rope-like strand called
‘sliver’. Yarn making requires that the fibers be well oriented and therefore must use
processes which will straighten and parallelize the fibers. An important one of these
is drawing, which is a process of elongating the strand to make the fibers slide over
one another and hence help in orientation.
Drawn sliver has to be reduced in size and then twisted to make yarn. There are
various processes of drawing (sometimes called drafting) and twisting. For some
specialized uses, yarns can be twisted together to form ply yarns.
Review of yarn production
15
1.4.7 Long-staple yarn production
There are three major types of long-staple yarn production systems; these are called
(a) worsted, (b) woolen, and (c) stretch-breaking. In practice, some worsted and
woolen manufacturing plants also make their own staple by cutting or stretch-breaking
tow. There is much more variation in plant layout for long-staple yarns than there is
for short-staple ones.
The worsted system was devised for twisted yarns made from wool but it has been
adapted for man-made fibers such as acrylic and its blends. In principle, it is similar
to the short-staple systems just described except that a wet scouring operation precedes
the mechanical processes, different types of card are used, and the machine elements
are generally larger than for short staple. It is an important process.
Woolen processing, despite its name, can also involve the processing of a variety
of long fibers ranging from wool to man-made fibers. The process (one of the original
short-process lines) consists of a card set comprising several roller-top cards assembled
in series on a single frame as well as spinning frames; and winders. The yarns are
softer and weaker than worsted yarns.
Stretch-break systems use only man-made filament tows and usually start with a
stretch-breaking machine, omitting the carding and preceding processes. It is a relatively
short process line. In carpet yarn production, the system is in competition with
bulked filament production, which has limited its growth as a system. Very occasionally,
the stretch-broken material is fed to a card when there is a desire to blend it with a
natural fiber. Process schedules for different products are given in Table 1.1.
1.4.8 Man-made carpet yarns
Carpet yarns are considerably heavier than apparel yarns and the poundage produced
is very large. Carpets produced with man-made yarns have displaced those made of
natural fibers in many markets. Nevertheless, the total market size has increased and
traditional carpets still hold a significant portion of it. This is of sufficient importance
to be mentioned at an early stage.
Several alternative manufacturing routes may be used, two of which are indicated
in Fig. 1.8. On the left is the traditional route in which tow is cut, oiled, and
crimped by the fiber maker before shipping bales of staple fiber to the yarn
manufacturer (the crimp is intended to make the fibers mutually cohere to make
further processing easier). On the right is a system in which the tow is shipped to
the yarn maker and the cutting or stretch-breaking is carried out in the mill to
convert the material to staple fibers in sliver form. The sliver is then bulked so that
the yarn made from it has a soft hand with good cover. Alternatively, the fiber
maker might bulk the yarn before shipping it directly to the carpet maker. The
productivities of the equipment in manufacturing bulked filament yarns are very
high and the economies of size favor the fiber maker over the small installations in
the mills in this particular field.
1.4.9 Composite yarns
A variety of systems are now emerging that combine filament and staple processing.
Some of these involve the wrapping of filaments round staple fiber bundles or low
twist staple yarns; some include a filament core for strength and others use binders.
These are referred to generally as composite yarns. Where low twist is used, the yarns
16
Handbook of yarn production
Tow production
Draw
Wind
Cut, oil
& crimp
Bale
Yarn
bulking
Yarn mfr
Bulking
Winding
Winding
Bulked yarn to end user
Fig. 1.8
Carpet yarn production
become softer in hand, but there is a danger that the filament will ‘grin through’3 the
staple fiber covering. This might result in an unacceptable appearance of the fabric.
Adequate cover of the filaments is generally required and this imposes limitations.
Other changes in yarn structure may also alter the aesthetics of fabrics and hence
much market research is needed for such new products.
The category of composite yarns also includes industrial yarns where the most
important attribute is strength. Some man-made filaments are exceedingly strong and
the technology of composite yarns becomes very important where there is a need for
a sheath of fibers with different characteristics that surround a strong core. A common
example is that of sewing thread, where a non-meltable sheath is desired.4
Some non-textile composite materials are made in which strong yarns are embedded
in a matrix to reinforce it. For example, concrete can be reinforced by high tenacity
yarns or fibers.
1.4.10 Review of processes
Since the cost of fibers is a large proportion of the final cost of yarn, it is important
that the spinner understands something of the processes and products of the fiber3 Grin through is a term used for exposure of the filament surface in the fabric, which may cause
changes in light reflection and areas of differing dye pick-up.
4 ln sewing, a thread passing through a needle at very high sewing speeds might be caused to melt
in places by the high temperatures in the needle eye. For example, nylon thread might melt on
the surface; but if it has a cotton sheath it will be protected because cotton does not melt.
Review of yarn production
17
producing industries. Not only do the average attributes of the fibers influence the
efficiency of mill processing and the quality of the yarn, but the defect levels and
variability within the fiber supply are also highly relevant. Thus quality control is
important. Consequently, this book may be regarded as containing three sections.
Chapter 1, which is a general overview, may be regarded as the first section. Chapter
2 is written in three segments to emphasize the materials employed and may be
regarded as the second section. The subsequent chapters deal in detail with the wide
range of processes used to convert fiber to yarn; these chapters may be regarded as
the third section of the book.
The subject matter covers a wide array of processes with many interlinked ideas.
There is an interplay of topics in the complex web of production processes and there
are some common principles that cut across the boundaries of the various process
sequences. For example, two important ones are drawing (or drafting) and twisting.
Drawing occurs in both filament and staple yarn processes. So does twisting.
Consequently, the next section begins with a discussion of these common principles,
which should lay a basis for the chapters that follow.
References
1.
2.
3.
4.
5.
6.
7.
8.
9.
Looney, F and Lord, P R. The Future is Not Just an Extension of the Past: Globalization –
Technological, Economic, and Environmental Imperatives, Textile Inst Ann Conf, Sept 1994.
Greenwood, F A. The Textile Mill of the Future: the Effect of Modern Technologies, Textile
Inst Ann Conf, 1973.
Cook, J C. Handbook of Textile Fibres, Merrow Publishing, Watford, UK, 1964.
The Wool Assoc of the NY Cotton Exchange, Inc. Wool and the Wool Trade, 2nd edn, Riverside
Press, Cambridge, USA 1955.
Morris D and Stogdon A. World Markets for Wool, The Economist Intelligence Unit, New York,
1996.
Brunnschweiler, D and Hearle, J W S. Tomorrow’s Ideas & Profits: Polyester, 50 years of
Achievement, The Textile Institute, Manchester, UK, 1993.
Livingstone, I. Cotton Inc, New York, US Cotton Textiles in a Global Environment, 5th EFS
Conf, 1992.
Livingstone, I. Cotton Inc, New York, Textile Competition in a Global Market, 7th EFS Conf,
1994.
Thomas, P R, Banfi, O, Brusadelli, E, Derencin, L, Gresteau, J P, Hansen, G, Hoffmeister, P,
Kampl, R and Leitner, J. World Markets for Spin Yarns: Forecasts to 2000, CIRFS Special
Report No 2637, The Economist Intelligence Unit, Dartford, May 1994.
2
Textile products and fiber production
Section A
2.1
Textile materials (fabrics, fibers, and filaments)
2.1.1 The nature of fabrics
Textile fabrics usually have the attributes of being soft and pliable with a capability
of being molded or draped over non-flat surfaces. The ‘hand’ of a fabric (which
describes its tactile characteristics) is very important in determining its acceptability
for many applications. For example, to obtain the desirable characteristics required of
apparel, the fabric must be made from fine yarns; also there must be some degree of
freedom for them to move within the fabric structure. The sensation obtained when
there is contact between human skin and fabric is determined to some extent by the
stiffness of the hairs or fiber loops projecting from the surface of the fabric. The finer
these outstanding hairs or fibers, the softer the fabric feels to the light touch. For this
reason, too, many fabrics are made from fine filaments or fibers.
2.1.2 Fibers and filaments
Let a filament be defined as a continuous fine strand whose length is so very long
that it can be considered infinitely long. Staple fibers, examples of which are cotton
and wool, exist in relatively short fiber lengths. We will refer to ‘continuous filaments’
and ‘staple fibers’ to help differentiate between filaments and fibers. Most natural
fibers are staple fibers although silk is an exception. (Silk is sometimes chopped to
make staple fibers.) Man-made fibers can be used as staple or continuous filament.
The very fine fibers used in textile fabrics make it necessary to use special units to
express the idea of fineness or ‘diameter’. Unfortunately, the many sectors of industry
have each invented their own systems over the years. Appendix 1 gives some of the
measuring systems used.
Textile products and fiber production 19
Another desired attribute of most textile fabrics is that they should have a good
appearance. This usually implies that the fabric must look even and have no blotchiness,
cloudiness, barré (see Fig. 2.1 and note that the bars are components of the condition
known as barré), or streakiness. This, in turn, implies that the yarns should have
uniform fineness, hairiness, and color along their whole lengths. With such materials,
faults such as thick and thin spots in the yarn and neps should be avoided because of
their disturbing effect on the appearance of the fabric. (Neps are tiny balls of fiber,
which degrade the appearance of fabrics). Also, light reflects differently from various
surfaces and a change in the structure of a yarn (or fabric) can cause unwanted
change in the appearance of the fabric. For example, variation in yarn twist, hairiness,
or fiber fineness can cause such undesirable changes. One must pay careful attention
to preserving quality in these respects; for this, and other reasons, quality control is
an important topic. Appearance is extremely important in the marketplace. With
many of the yarns, great efforts are made to ensure yarn uniformity in all details.
However, there is a special class of fabrics, which uses color patterns, random
disturbances (such as nubs), and thick and thin spots in the yarns to produce effects
which give interesting textures; such yarns are termed ‘fancy yarns’ or ‘effect yarns’.
Nubs are random thick spots induced into the yarn to produce visual effects in the
fabric.
Each sort of yarn has its set of physical and mechanical properties which influence
the performance of the yarn. To best select yarn for a particular purpose, one must
understand the properties of the materials. Since there is an interaction between the
technology and the properties of the yarn produced, it is very important to study all
aspects of these topics.
2.1.3 Classification of fibers
The principal division is between natural and man-made fibers. The ‘natural’
classification is subdivided into fibers of vegetable, mineral, and animal origins.
The vegetable subclassification is shown partially in Fig. 2.2. Dotted lines indicate
that there are other members of this particular subdivision that need not be considered
here because of their small market share. Cotton fibers are the most important in this
category, and they will be considered. Stem fibers are also known as bast fibers.
Details of the fibers in the two bottom rows will be described later. Full classifications
can be found in the literature [1–4]. About the only mineral fibers are asbestos and
glass. Asbestos, in modern times, has been associated with asbestosis and has been
banned in many parts of the world. Glass has a substantial market in industrial
insulation, non-wovens and, of late, optical fibers for communication. The glass is
Bar 1
Bar 2
Fig. 2.1 Knitted fabric with barré
20
Handbook of yarn production
Natural fibers
Vegetable
Stem
Leaf
Mineral
Seed
Cotton
Flax
Fig. 2.2
Jute
Hemp
Animal
Hair
Filament
Wool
Silk
Natural fibers – partial classification
melted and extruded and the techniques of production have much in common with manmade fibers. Wool is the most important animal fiber and large quantities of fiber have
been used for carpets and apparel. Like the vegetable ones, they are natural and variable.
2.1.4 Polymeric materials
Fibers are made of polymers; some are natural polymers and some are man-made. A
class residing in between these two consists of regenerated fibers, made, for example,
from regenerated cellulose from trees, waste cotton fibers or others of the many
natural sources of cellulose, and modified natural fibers, made by reacting natural
polymers with chemicals to alter their properties. It is a misconception to view only
synthetic fibers as polymers. Natural polymers are cultivated in the agricultural sector
whereas man-made fibers are produced in the industrial sector. The synthetic polymer
is first produced in chip or similar form, or is supplied directly to extruders in a liquid
state. If it is produced in chip form it is later melted or otherwise liquefied and extruded.
A textile polymer is made up of long-chain molecules. A long-chain molecule may
be regarded as a long string of atoms; these ‘molecular strings’ are flexible (if they
are not cross-linked) and give to the fibers many of their desirable characteristics.
The analogy between the behavior of a long flexible fiber and a long flexible molecule
is no accident. However, the analogy must not be carried too far because there can be
strong bonds between the molecular chains that make up the fibers, which are not
simulated by the fibers in a yarn.
The polymer has to withstand the end use conditions. For example, it would be of
little use making a fabric that would melt or soften in hot water. It also has to be
strong enough to fulfill its purpose. Other properties have similarly to be taken into
account.
Many polymers can be set by raising their temperature above the so-called glass
transition temperature (Tg ), deformed and then cooled. Tg is the temperature at which
the polymer softens. Some fibers (such as cotton) cannot be permanently heat set, but
easy care properties can be induced into fabrics by a chemical treatment called cross-
Textile products and fiber production 21
linking. This treatment joins groups of molecular chains together and reduces their
ability to move with respect to one another. The linking reduces the loss of energy of
deformation and makes it more likely that the retained energy is available to return
the fabric to its original shape. The cross-linking also makes the fiber structure
stiffer. Many cross-linked fabrics, especially for apparel, have easy care properties.
Some fabrics have creases or shaping set into them; thus, even after laundering or
cleaning, they retain the desired shape or crease. One class of yarns is made by
setting the filaments themselves into certain shapes and this is called texturing. When
heat or mechanical stress causes a variation in the molecular structure, it can alter the
way dye is taken up at various places along the length of the yarn. It can cause a
fabric to look streaky.
A classification diagram for some man-made fibers is given in Fig. 2.3; again,
some members of the various subdivisions are omitted. Polyesters’ long-chain synthetic
molecules are composed of esters of aromatic dicarboxylic acids and glycols. There
is a family of polyesters but one of the most common is polyethylene glycol terepthalate
(PET), which is widely used in staple form. It is often blended with cotton for
apparel. The blends give some of the benefits of each component. The moisture
absorption and feel of cotton are perceived to add comfort to apparel made of the
fiber. Polyester has durability and recovery properties that add to the easy care
attributes of any fabric made from it.
Polyamides are long-chain synthetic polymers made from diamines and dicarboxylic
acids. The most widely used are collectively known as nylons and the various chemical
types are indicated by adding numbers which indicate the monomers from which they
are formed, e.g. nylon 6, nylon 6.6, and nylon 11. They are widely used in both staple
and filament forms for carpets. Growth of this market has been at the expense of
wool. Acrylic fibers are another class of synthetic polymer, composed of at least 85%
by weight of acrylonitrile units. Where less than 85% and more than 35% of the
polymer comprises acrylonitrile units, the fibers are termed ‘modacrylics’. Textured
Man-made fibers
Regenerated
Synthetic
Rayon
Polyamide
Acrylic
Polyester
Fig. 2.3
Polyolefine
Man-made fibers – partial classification
22
Handbook of yarn production
acrylic fibers have become popular for garments such as sweaters and have displaced
wool for some portions of the market
Polyolefines, such as polyethylene and polypropylene, are made by polymerizing
olefins such as ethylene (ethene) and propylene (propene). Polyolefines have come
into widespread use for wrappings and have displaced jute from a large segment of
that market. Wrapping fabrics are not necessarily made from conventional yarns;
they may be non-wovens, or made from tapes rather than the more or less cylindrical
yarns. These alternatives will not be further discussed. There are a number of special
polymers used to make high performance fibers and filaments for industrial applications.
For example, various aramid fibers have high tenacity and high temperature resistance.
They are a form of polyamide. There are also the polyurethanes, some of which are
endowed with enormous elongational capabilities.
2.1.5 Staple versus filament
As previously mentioned, most natural fibers exist in discrete lengths whereas manmade fibers are produced as extremely long filaments, which can be processed as
continuous filament yarn or may be converted into staple fiber. Natural fibers are
agricultural products subject to changes in properties due to exigencies arising from
variations in growing conditions. Man-made fibers are usually controlled more closely
but variations still occur; many of the factors that do vary are rather subtle in
their effects. There exists an extremely important demand from the non-technical
end user in favor of the natural fibers and blends between natural and man-made
staple fibers. Aspects of hand and appearance in consumer products are very
important in this market and some of the expectations of better quality from manmade fibers arising from closer control have not been realized. Users of technical
products such as ropes, belting, and other industrial materials are usually more
concerned with strength rather than appearance and technical factors assume a greater
importance. The point is that there are market divisions with widely differing
requirements.
The production of the two classes of fiber, natural and synthetic, is radically
different and has to be discussed separately. Since natural fibers are those with the
longest history, they will be discussed first. The sharp contrast between the methods
involved in the two cases will be noted. But it will be appreciated that often a spinner
has to deal with both man-made and natural fibers and there is a need to know about
the sources and idiosyncrasies of both.
Section B
2.2
Natural fibers (types and production)
2.2.1 Cotton
Cotton fibers
Cotton is a vegetable fiber that grows from the surface of the seed. Each fiber is
essentially a long thin tube of cellulose with a central feed channel, called a lumen,
Textile products and fiber production 23
which runs almost the whole length of the fiber. In modern production, cotton is
cultivated as an annual plant rather than letting it grow into a tree. Harvesting is
easier working in this way and fiber properties are better controlled; also, cotton
plants left in the ground after harvesting are subject to attack by pests. The use of this
so-called ‘stump cotton’ has been banned in the USA.
The length to diameter ratio of the fiber is in the order of several thousand; this
makes it mechanically flexible and suitable for textiles. Wild and cultivated species
of cotton [5] have been placed in the genus of Gossypium and in the order of Malvales.
Five species have been cultivated, Gossypium herbaceum, Gossypium arboreum,
Gossypium hirsutum, Gossypium barbadense, and Gossypium peruvian. The first
two of these are sometimes known as Asiatic species, the third is commonly known
as American upland cotton, and the last is known as long-staple cotton (Sea Island,
Egyptian, Peruvian tanguis, Pima and others). Asiatic species have historically been
known for having short fiber lengths, whereas the G. barbadense is known for longer
fiber length than the average. Much breeding work has been done to change the
characteristics of the short fibers, particularly if they were coarse (i.e. of high linear
density).
Cotton fibers are elongated single plant cells, varying in length, the average length
of which changes according to species and conditions of growth. Fibers develop as
elongations of the outer layer of cells of the cotton seed and each fiber consists of
layers. The outer and inner layers are called the primary and secondary walls, respectively.
The wall has a structure of fibrils as sketched in Fig. 2.4(a). A growing fiber exists
as a tube with a wall thickness defined by the primary and secondary walls. The wall
thickness and length alter as the fiber grows but there is little change in the outer
diameter of the fiber. A cross-section of the undried fiber reveals the existence of
daily growth rings, in more or less concentric circles surrounding the lumen. At
temperatures less than 72°F or more than 90°F, the plant becomes dormant but within
this range, growth occurs and the diurnal changes in temperature produce the growth
rings.
An erstwhile fiber tube flattens when it matures and dries. The deformed tubular
fiber gives a variable cross section that is sometimes as sketched in Fig. 2.4(c) and
sometimes in other configurations of a flattened tube as shown in the micrograph in
Fig. 2.4(d). According to Scott Tagart [6], the wall thickness is not constant, and
when the lumen dissipates, an irregular collapse is caused by irregularities in the
wall. It is not possible to show the length of the fiber in proportion to its cross-section
and only a part is shown in Fig. 2.4(b). The convolutions or spirals of the twisted
ribbon of the dried fiber make it easily spinnable, which is an important consideration.
It has been stated that there can be as many as 250 twists along a single fiber but the
direction of twist does not remain constant; there are a number of reversals along the
length.
Wall thickness is also important because immature fibers with thin walls tend to
collapse into neps during processing. These neps are a great nuisance during processing
and can seriously degrade the value of the final product. Also wall thickness and
structure can affect dye affinity which, in turn, affects the color of the finished
product. Fiber length is also of great importance and a premium is paid for long
fibers.
Extremes in length vary from less than 0.5 to over 2 inches. The former is of little
use in yarn manufacture and the latter is expensive and somewhat rare. Also various
species of cotton have differing average diameters when growing and, of course, the
24
Handbook of yarn production
Fibrils
Wall
Lumen
(a)
Mature dried fiber
(b)
Cross-section
(c)
(d)
Fig. 2.4
Cotton fiber
diameter varies with the growing conditions. The concept of diameter is of little value
because of the changes caused by drying as just mentioned. Rather, it is preferred to
talk of fiber fineness or linear density.1 Color of the fiber is also important; some
species are white and others have various depths of yellowness, which affect the final
product. There are some cotton fibers which have been bred to have a range of natural
colors but these only occupy a small proportion of the market at the time of writing.
Variations in fiber characteristics make it very important for a mill to choose the
fiber carefully and to blend it well and consistently.
After the cotton flower has gone, a ‘boll’ is formed, which is a fairly small, fat,
pear-shaped capsule. It bursts open as it ripens exposing the mass of fluffy white
fibers. Seed fuzz begins to develop from epidermal cells at about one week after the
opening of the flowers, whereas the fibers useful in spinning begin to form on the day
of flowering. The short fuzz is known as ‘linters’ and is sometimes used for paper
making, wadding or as a source of cellulose. Fibers useful for spinning are known
collectively as ‘lint’; these useful cotton fibers are obtained from the output of the
gin.
1 Mass/unit length of fiber. Also, the term ‘micronaire’ is used which is really a standardized test
of air permeability of a given mass of fiber in a certain restraining volume.
Textile products and fiber production 25
Fibers are attached to the seed at one end of the fiber, and they develop as the seed
ripens. The fiber, at the attachment point, is hooked, and this portion of the fiber is
susceptible to nep production. The aggressiveness of the ginning process (separating
of fibers from unwanted material) determines how much of the hooked portion
passes into the lint used by the spinner and thus the nep potential of the fiber.
There is continuous activity by breeders who strive to improve the fiber length,
strength, and other fiber properties. For example, over the period from 1990 to 1994,
the average fiber strength of US cottons improved from 26.3 to 28.4 gf/tex. The
strongest specimens yielded values as high as nearly 40 gf/tex. Fiber fineness is also
judged to be very important. In addition, the breeders strive to introduce varieties
which are resistant to disease and give high yields. In recent times, genetic splicing
[7] is augmenting the practice of culling the best from a large number of varietal
developments using traditional techniques.
Cotton growing
Cotton is a tropical plant requiring moisture and sunshine. Cultivation requires a
climate that avoids frost damage to the growing plant. If the climate supplies insufficient
rainfall, then it is necessary to use irrigation. The requirement for sunshine means it
is normally grown between the latitudes 47°N and 40°S, which includes the Americas,
Africa, Australasia, and Asia. Little is grown in Europe.
More than 500 species of insect attack cotton plants and many of them are very
destructive. Lush foliage, large flowers with their nectaries, and the extended fruiting
period make cotton a host for many insects. Consequently, care has to be taken to
control boll weevil, bollworm, aphids, nematodes, and other pests, but the use of
modern pesticides has reduced the problem. The boll weevil and such insects destroy
the boll by consuming leaves, bulbs, and the bark of the plant. Damage caused by
aphids, whitefly, and the like is more subtle. Secretions deposited after the boll is
opened convert to sugar [8], which makes the fibers sticky and difficult to handle in
processing. Deposits on metal surfaces in ginning and mill equipment can cause
substantial losses in production. Control in the field is not simple. Early season
pesticide application can disrupt the natural suppression of other pests and accelerate
resistance to evolution. Mid-season aphid populations cause yield reductions and late
season populations produce sticky cotton. Some aphids are highly reproductive under
the optimal environmental conditions. The presence of other aphid hosts such as
sweet potato nearby can increase the damage because of the increased populations
surrounding the cotton fields. For example, cotton leaf hair is an important factor in
infestation of cotton [9] by the Bemesi tabaci whiteflies from the sweet potato. The
infestation produces sticky cotton. Sugar deposits can come from other sources, such
as honeydew (a secretion from aphids). Irrespective of source, they pose a problem
that still causes anxiety. Pathogenic fungi, bacteria, and viruses also attack cotton.
The point of this discussion is to show some of the difficulties and underline the need
for mills to watch for these deposits and infestations because of the difficulties they
can create in yarn manufacture. Various methods of measuring cotton stickiness [10,
11] are available.
Yield, disease resistance, fiber length, strength, fineness, and maturity are primary
factors used to select cultivars [12].2 As time goes by, new strains of cotton are
2 A cultivar is a plant of a kind which originates and persists under cultivation.
26
Handbook of yarn production
produced which improve the performance of ginning and textile machinery, as well
as enhance the value and quality of the product. The growing of cotton needs skill and
attention but space precludes any great discussion of matters other than those which
affect the spinner. Suffice it to say that the rows of plants have to be spaced to allow
agricultural machinery to move down the rows. Modern practice uses picking, spraying,
and other machines that straddle two or more rows. Because of this, the machines are
characterized by their tallness. There is a necessity to control weeds, especially those
which produce oily seeds, act as hosts for the pests just described, or otherwise inhibit
growth of the fiber.
Before harvesting, the plants are defoliated by chemical means; much of the
embrittled foliage falls off in the field and some is removed in handling. This reduces
the trash in the fiber before it is harvested. In advanced countries, bolls of fiber are
machine-picked and transported to a gin in a module (which contains about 6000 lb
of cotton) or trailer (which contains, on average, about 4000 lb).
Two types of mechanical harvesting involve the use of a spindle picker and a
cotton stripper. The processes are often referred to as ‘machine picking’ and ‘machine
stripping’ respectively. The spindle picker uses tapered, barbed, rotating spindles to
remove seed cotton only from well-opened bolls. ‘Seed cotton’ refers to material
from which the seed has not yet been removed. The cotton stripper is non-selective;
it includes cracked and unopened bolls along with burrs and other foreign matter in
its output. Fouled spindle pickers produce spindle twists that are twisted masses of
fiber difficult to handle in any process. Stick content and grade reduction due to bark
is controlled by the aggressiveness of the roll settings on the stripper. Under conditions
of improper setting and wear, particles of rubber from the stripper become embedded
in the lint and cause ‘black spotting’ in the yarns ultimately produced. Thus, the
product reaching the gin is affected not only by the seed and the growing conditions,
but also by the setting and maintenance of the harvesting machinery. Unopened
(‘green’) bolls are not only regarded as non-lint but they also carry considerable
moisture. When the green bolls burst, the moisture migrates through the surrounding
seed cotton. Changes in retained moisture in the supply can upset the moisture
content of the cotton already being processed; defects may be created in the product
supplied to the mills. Moisture content is frequently expressed as ‘moisture regain’,
the amount of moisture in a sample compared with that in an oven-dried sample. A
fairly typical distribution of unwanted material in lint is given in Table 2.1.
As the growing plants pass their harvest time, they are prone to become gray
because of weathering. The decision when to pick can provide a problem if the
weather is uncertain. Harvested fibers left out in the open can become discolored and
stained by exposure to the weather; the severity of this staining is thought to be
Table 2.1 Non-lint content of some cottons. Values expressed as
mass percentage, i.e. {(non-lint/useful lint) × 100%}
Burrs
Sticks
Fine trash
Motes
Total non-lint
Machine
picked
Machine
stripped
7%
2%
5%
6%
20%
90%
23%
22%
5%
140%
Textile products and fiber production 27
influenced by the amount of wax produced by the plant. Thus, differences between
varieties and growing sites can play a part in this respect. However, moisture content
is a most important factor and length of storage is obviously a factor too. Consequently,
care is taken to keep the newly harvested cotton under proper storage conditions.
Suitable outside storage sites [13] should be well drained and free of gravel, stalks,
long grass, and debris. They should have a smooth, firm, and flat accessible storage
surface. Obviously staining or weathering can degrade the part of the crop affected.
The variability of the products in the various stages of yarn manufacture will be
discussed in later chapters. However, the stage is set by the variabilities within the
cotton itself. For example, it has been variously reported that the fiber fineness
within a single plant can have coefficients of variation of the order of 15%. This can
be compared to coefficients within a single field of cotton that are of the order of
10%. These variations become partly homogenized by the blending that occurs in the
various process stages but it is important to realize the magnitude importance of
these sources of variance.
Cotton ginning
An important step in the production sequence is ginning. In the ginning operation,
sticks and coarse trash are removed from the input material. Also, since the input to
a normal gin is from a variety of growing areas and a gin processes very large
quantities of fiber, the process acts as a first blending of massive quantities of fiber.
The term ‘gin’ is sometimes used to describe a whole establishment that provides the
service of ginning, but sometimes it is used to describe the machines. The part of the
machine line that separates fiber from seeds is often referred to as a gin stand.
The Churka gin, which, like all gins, had the purpose of separating the seed from
the lint, was developed in some unknown period BC; Eli Whitney is credited with
inventing the cotton saw gin in 1794. The basic idea was, and still is, to grip the lint
protruding through a set of ribs by a moving surface and wrench the fiber from the
seeds, which are unable to pass through the ribs. Rotary saw blades are commonly
used as the moving surface just mentioned, but a specially prepared roller surface is
sometimes substituted. A perfect operation would result in damage to neither seed
nor fiber, and the undesirable small portions of fiber at the attachment points to the
seed would be excluded. Unfortunately, this state of perfection is not possible. The
aim is therefore to minimize the damage and thereby maintain the salability of both
products. Apart from this, ginning has developed into a process stage that involves
more than just removal of the seed from the lint. Before going on to discuss the
elements in ginning, let us consider the seed.
Seed is an important by-product of cotton manufacture and most of the seed is
crushed for oil or used for animal feed. The percentage of the US crop that was
crushed for oil declined from over 85% in 1970 to about 50% in 1994. On the other
hand, the percentage used for feed changed from about 10% to over 35% over the
same time. To the spinner, the seed represents a hazard because, if seed-coat fragments
are excessive, mill processing is made more difficult. The economics of ginning are
affected by the sale of the by-products, and thus they have some effect on the cost of
cotton to the mill. Ginning, warehousing, and merchandising [14] accounted for 90%
of the cost of cotton lint in 1977, of which over half was taken up in warehousing and
merchandising. The costs generated between the ginning and the mill processes are
significant. It becomes all the more important to ensure that, not only is the cost
controlled, but also that the quality of the product is maintained within specified
28
Handbook of yarn production
limits. The cost of cotton to the mill represents roughly half that of yarn and the
spinner must therefore have a strong interest in the basic fiber production.
A modern ginning establishment [15] contains some or all of the following process
items:
1
2
3
4
5
6
7
8
9
10
One or more green boll traps, used to remove green bolls, and heavy foreign
matter such as rocks. These green boll traps are often little more than a Ushaped diversion to the flow of air and seed cotton. A counter-weighted trap
door releases the green bolls when they have collected in sufficient quantity to
counterbalance the forces holding the trapdoor shut.
Automatic feed controls to provide an evenly dispersed fiber flow without wet
clumps.
Dryers such as is shown in Fig. 2.5 – two dryers are normally used which
expose the flowing fibers for 10 to 15 seconds to hot air and raise the fiber
temperature to no more than 350°F.
Two or more cylinder cleaners – these consist of spiked cylinders rotating at
400 to 500 r/min over grid bars. Interaction between the moving and fixed
elements breaks up large wads of fibers to permit more even distribution of
moisture and temperature among the fibers. This, in turn, induces removal of
fine foreign materials such as leaves, trash, and dirt.
‘Stick’ machines to remove burrs and sticks.
A conveyor-distributor to convey cotton to the gin stands, where the separation
of fibers from seeds takes place.
A feeder to control the flow rates to the gin stands.
The gin stand – this is the central item in the system, about which more is
written in the following paragraphs.
Lint cleaners to remove immature seed fragments and other trash.
Bale presses to compress the lint into bales to facilitate transport and storage.
The need to clean the fiber at ginning is driven by the urge to elevate the grade to
get the best possible financial return for the farmer. However, a high grading obtained
by excessive cleaning always causes disappointment in the mill due to fiber degradation
and the fiber may, therefore, not fetch the high price envisaged.
Hot air + cotton
Dried cotton
Fig. 2.5
Drying fiber at the gin
Textile products and fiber production 29
To make removal of trash easier, the cotton is heated to get the moisture content
down to about 7%. Hot air is used for the drying, and the time to dry the cotton
depends on the starting moisture content and the air temperature and flow rate. (The
moistures content of a fiber in hot air declines approximately exponentially with
time.) There is a temptation to use high air temperatures in order to speed up the
process. However, this, or the slowing of the fiber flow, or changes in initial moisture
content can lead to overheating. Associated ills are damage to the fibers by scorching,
electrostatic charging of the fibers, and driving of the more volatile components of
the wax that coats the fibers. Wax on the fibers is a valuable lubricant without which
later processing would be very difficult. Electrostatic charges impede the separation
of the fibers and lead to fiber damage. For those reasons, too low a moisture content
arising from overheating the fiber is counterproductive.
A modern saw gin stand [16] has between 100 and 200 disc-like saws separated by
ribs about 0.5 inch wide. Saws up to 18 inches in diameter are used. A simplified
sketch of a typical saw gin stand arrangement is given in Fig. 2.6(a), but the sizes of
the saw-teeth and brushes have been exaggerated for clarity. The cutting edges of the
teeth are angled (line CD, which is parallel to the tangent of the rib AB) as shown in
Fig. 2.6(b). The reason for this is to prevent the seeds from sliding to the base of the
teeth and accumulating there. The gin stand comprises a feed system and spiked
Seed cotton
Brushes
Saw
Lint
Gin rib
Huller rib
(a)
Debris
Rib
(b)
Fig. 2.6
Elements of a saw gin stand
30
Handbook of yarn production
roller cleaners, as well as the gin components shown. Seed cotton is introduced to the
saws by a feed roller. The teeth of the saws pull the lint through the rib gaps into a
moting chamber (not shown) where seed particles are removed by a mote bar. From
there, the lint is carried to the brush doffer, which removes the lint and transfers it to
the air transport system which, in turn, carries the fiber to the lint cleaner. Most
cleaners consist of saws and mote bars that flail the lint against the bars. Brushes
move lint from the saws.
An alternative to the saw gin is the modern roller gin (Fig. 2.7) which was developed
in the 1950s and is used mostly in Arizona, California, Texas, and New Mexico. An
important feature of this gin is the covering of the roller, which has to be resilient, yet
durable, and have a surface that will grip the fiber. The productivity is only about
20% that of a saw gin. Since ginning accounts for about 38% of the cost of cotton,
roller ginning is used only when there is a particular need to preserve fiber quality.
Pima cotton (Gossypium barbadense) and similar long-staple cottons are usually
roller ginned.
Woody and other unwanted materials are unavoidably collected in the fiber when
mechanical harvesting has been used. All unwanted matter is collectively referred to
as trash. Seeds and non-lint materials are removed in cleaning but some good fiber
is also lost. Too rigorous a cleaning can damage fibers and degrade the product. In
seed removal, it is inevitable that some fiber damage will occur; seed particles are
removed by the violent action of separation. Moisture is very important in this respect.
For example, the strength of upland cottons is about 1.8 times the force needed to
separate the fiber from the seed at 7% moisture content. The short fiber content of the
lint is increased by about 1% for each percentage reduction in moisture content [17]
This is partly because the increased fiber electrification in the ginning process makes
fiber separation more difficult. Some unwanted matter is taken out by the lint cleaners,
which operate in series with the gin. It is a matter of debate how much cleaning
should be done at the gin and how much in the mill. If dirty cotton is shipped to the
mill, the yield of usable cotton per bale drops and the effective cost of the fiber rises,
but the quality of the fiber is better. Similarly, the mechanical removal of fiber, which
accompanies any cleaning operation, reduces the ‘out-turn’ of the gin. (Out-turn is
defined as the mass output as a percentage of the corresponding input.) This tends to
reduce the financial return to the ginner and farmer, unless there is a compensating
Seed cotton
Moisture
spray
Rotating
knife
Stationary
knife
Ginning
roller
Debris
Lint
Fig. 2.7
Elements of a roller gin
Textile products and fiber production 31
increase in premium due to improved cleanliness or grade. The use of a larger number
of lint cleaners leads to a reduction in trash content but only at the expense of a
decrease in fiber quality. The addition of a third lint cleaner in ginning causes a
greater loss in fiber quality than that caused by the second cleaner. Since fiber
damage is irreversible, there is some question as to the wisdom of using a third lint
cleaner in ginning. However, the question is one that is usually settled by the market.
After the completion of ginning, the fibers are compressed into bales to facilitate
transport to the mills. A bale usually weighs about 500 lb and is wrapped in a fabric
to protect it from damage, and strapped with wires or metal tapes to maintain the
compression during transport and storage. These ties are strained up to 2000 lb force
and extreme care has to be taken in bale breaking (i.e. removing them). The pressure
used to compress a bale is also of importance. If several bales decompress to differing
heights when introduced into the bale laydown in the mill, the bale plucking machinery
will not sample the bales correctly until the top of the laydown has been leveled.
Improper sampling of this sort could lead to barré (see Fig. 2.1) in the final product.
Bales come in a variety of sizes and, even within the USA, there are several standards
such as flat, modified flat, and universal. A bale is box-shaped (i.e. a rectangular
prism) whose dimensions are X, Y, and Z. With US cotton, X ≈ 55 inches, Y ≈ 20, 21,
25, or 28 inches, and Z depends on the degree of compacting. For a 500 lb bale,
densities vary between 20 and 30 lb/ft3. In earlier times, bast fibers were used to
make the bale wrapping, but now they are commonly made from polypropylene tape.
The wrappings assume significance because failure to remove all vestiges of the
wrapping leaves ‘foreign fibers’ in the product stream, which cause blemishes in the
final product.
2.2.2 Wool
Wool fibers
Wool is normally defined as the fleece of sheep, the fleeces from other animals can
also be used as textile fibers. Camel and goat hair (some goat fiber is known as
mohair) are highly prized although they have only a small market. For this reason,
they will not be described here.
Wool is a protein called keratin, which has a main polypeptide chain with amino
acid side chains. It is an outgrowth of the epidermis (skin) of the sheep and the
surface of the fiber has minute overlapping scales extending lengthwise and pointing
to the end remote from the root or cuticle. The root is embedded in the epidermis.
Wool grows in tufts, in or near the follicles in the skin of the animal; however, the
useful, outermost portions of the fibers on the animal are no longer growing. Growth
occurs by multiplication of the soft cells of the papilla, which exist at the base of the
follicle. The papilla is a vascular arrangement of connective tissue extending into and
nourishing the root of a hair. The useful part of the fiber is displaced from the cuticle
as new cells are added and the fiber gets longer. A scaly surface is produced, as
shown in Fig. 2.8. Thus, the fiber grows in length even though the outer part is no
longer living. Lack of nutrition, or disease, affects the development of the fiber; if
there are periods when the animal is adversely affected, portions of fiber become
weak. Wool with such weak spots is referred to as ‘tender’. The central part of the
fiber near the skin contains the medulla (the inner pithy part of the structure) and the
cortex (a layer of tissue connected to the papillae).
Certain moth and beetle larvae such as Tinea bisselliella, Tinea pellionella, and
32
Handbook of yarn production
Fig. 2.8
Portion of wool fiber showing scales
Anthrenus verbasci attack wool and it is necessary to protect the fiber. Mothproofing
does not always protect against beetle larvae because the eggs may have been laid
before the treatment. Some insecticides do not protect against all pests that attack the
fiber and insects build up resistance [18]. Thus, it is desirable to have a broad range
of protection. Common agents used are organochlorine compounds; microbial pathogens,
parasites, and biological control also offer possibilities of control.
Wool is a source of anthrax contamination in humans but heating during the
drying cycle is sufficient to be lethal to the micro-organisms concerned. Any danger
comes from handling untreated wool from geographic areas having problems with
the disease.
Coarse wools are stiffer than finer wools and this property is especially important
at the tips of the fiber. Garnsworthy et al. [19] deduced that, if the fiber ends protruding
from an apparel fabric are capable of supporting loads of over about 1 mN, the wearer
will suffer mechanical irritation of the skin. These stiff fiber ends irritate the human
skin by mechanical disturbance of the surface of the epidermis. The effect can be
lessened through surface treatment of the fabric, but the main point here is that
fineness of the fiber is a significant property. In some applications, such as in carpeting,
stiffer fibers are preferred, but in most, more flexible fibers are preferred to avoid the
problems of ‘prickle’.
Follicles are deep pits in the skin of the animal where the fibers are nourished and
where wax and sweat glands are situated. (Sweat glands are known as ‘suint glands’.)
Thus, wool removed from the sheep is coated with wax and suint, as well as vegetable
matter picked up by the animal during foraging. Raw wool in this state is referred to
as ‘greasy wool’ or ‘grease wool’. The foreign materials, which can represent over
half of the weight of the unprocessed fleece, have to be removed before the normal
mechanical textile operations can be performed. Such cleaning operations used to be
carried out in the mill but, increasingly, they are being performed nearer the shearing
operation. Cleaning is divided into wet and dry cleaning; the former operation is
referred to as ‘scouring’. Further discussion of cleaning is given in Chapter 8.
Wool fibers are naturally crimped. The fiber crimp levels range from 6 per inch (in
cross-bred wools) to about 24 per inch (in Merinos). The fiber cross-section is slightly
elliptical and the balance of forces in the growing fiber cause it to curl in growth. The
fiber crimps are related to this curling.
Sheep bred for fine wool usually produce almost white wool and the presence of
any color can downgrade it. The color can vary from white through yellow and fawn
to brown; also shades of gray may be present.
Tenacity varies in the approximate range of 100 to 150 mN/tex when dry but only
80% to 90% of that when wet. Elongation varies between about 25% and 35%. The
average diameter of the fibers varies from 8 to 60 microns (approx 0.0003 to 0.0024
inches) and the useful length can be anywhere between a fraction of an inch to 40
Textile products and fiber production 33
inches. Obviously, the length depends on the growing time, the breed, and other
considerations. Some fibers consist of only cuticle (outer skin) and cortex whereas
others are medulated; some fibers contain parts that are medulated and parts that are
not. The term ‘medulated’ refers to a pithy core sometimes found in a fiber. Besides
normal wool, the animal grows ‘kempts’, which consist of coarse straight fibers with
tapered tips that are often shiny. They show up as undesirable inclusions in the final
product and therefore should be avoided. Discriminating between the various sorts of
wool is the job of the ‘classer’ (classifier) and it will be realized that classing is a
highly skilled job; much of the skill is based on experience.
No single, universal classification system exists, but one commonly used grades
the fiber according to the finest yarn that can be spun with it. An alternative system
is the ‘blood’ system, which refers to the closeness to the Merino breed of sheep.
Highly graded fibers are classified near to 100% blood and, as the percentage blood
reduces, so the quality is lowered. Another system grades from ‘super’ down through
AAA, AA, A, and 1st, to 2nd. As we have seen elsewhere, fiber length, fineness,
strength, elongation, uniformity, color, and luster are important and are all parts of
the value judgment. Fiber lengths greater than about 2.5 inches are regarded as
suitable for combing and for processing into worsted yarns. Fibers less than about
1.25 inches are used in the woolen system. The yarn count is usually quoted in the
woolen or worsted systems, where the count is measured in hanks/lb. The lengths of
yarn in these types of hank are 1600 yd and 560 yd, respectively.
Wool production
Wool is obtained by shearing it from living sheep or by pulling it from the skins of
slaughtered sheep. Shorn or sheared wool is the most common. Pulled wool may
alternatively be described as ‘skin wool’, or ‘slipes’. Fiber taken from a dead sheep
is called ‘dead’ or ‘murrain’ wool [20].
Australia and New Zealand produce a large percentage of the world’s wool. Much
of New Zealand wool [21] comes from the NZ Romney sheep and its crosses; Coopworth
and Perendale breeds are also significant. In Australia, the finer Merino wools
predominate. In these countries, very large areas are often involved in sheep raising.
Sheep can exist in rough mountainous regions under arid conditions, where forage is
variable and herbage sparse. With large flocks, careful organization is required to
harvest the wool because it is necessary to collect and confine the animals in close
quarters to permit marshaling to allow the clip to proceed efficiently. Fine wool
Merino sheep are able to survive under such conditions. Naturally sheep can also do
well with lusher vegetation, as evidenced in England and New Zealand. However,
some of the wools produced are relatively coarse and more suitable for carpets than
apparel.
Each sheep is periodically sheared to remove the fleece. The natural mutual cohesion
of the fibers enables them to cling together even after shearing. A skilled shearer clips
close to the body to keep the fleece essentially in one piece, and produces over 100
fleeces per day. Removal of heavily soiled wool around the crotch (crutch) of the
animal is called ‘crutching’ and it is usually carried out before lambing and before the
normal shearing; this reduces the amount of soiled wool to be removed later. After
shearing, the fleece is sorted and classed. Sometimes the fleece is skirted locally (i.e.
inferior pieces such as the short belly fiber are removed) and these skirtings are
packed separately. Dirty or stained wool and rough, coarse ends are removed. Sweaty
locks and parts containing a high percentage of burrs are also removed. All sorts of
34
Handbook of yarn production
vegetable matter picked up by the animal in its wanderings are usually included in the
term ‘burrs’. The remaining fleece is divided and rolled to bring the rib and shoulder
portions to view. The fleeces are then classed and shipped to warehouses for storage
and eventual distribution. Discolored wool, ‘bellies’, and some other portions are
sometimes sold separately from the main fleeces. Also wools containing faults are
separated; these include cotted, stained, and muddied wool, as well as double fleeces.
The term ‘cotted’ refers to heavily felted and matted material; double fleeces refers
to the material removed from sheep missed in the previous shearing. Annual shearing
is the normal practice and the yield thus normally depends on the growth rate of the
wool. If left longer than a year, the fiber length may be more than that obtained with
present practice; however, any increase in value with fiber length has to be set against
the reduction in average annual yield.
The jobs described are seasonal and are highly labor intensive; this drives up the
cost of production. Some consideration has been given to dosing the animal with a
drug to weaken the fiber near the follicle [22] so that the wool could be pulled from
the skin surface rather than be sheared; it will be interesting to hear of the outcome.
As mentioned previously, some movement has been made to clean the fleeces before
sending them to the mills. Scoured wool and ‘tops’ command higher prices of course;
the cost of shipping is lowered and environmental problems for the buyer are greatly
reduced. ‘Tops’ are thick ropes of fiber, otherwise called sliver. Some 65% of New
Zealand wool is now scoured before shipping. In contrast, as of 1988, 80% of the
total Australian clip was shipped as greasy wool. However, it should be noted that
wools from these countries go to different markets because of the differences in
fineness of the wool and the differing preferences of their buyers.
Arriving fleeces are closely examined by the buyer, despite the prior classing.
Different qualities, known as matchings, are placed in separate lots to maintain
quality control. The necessity to re-class underlines the variability of the wool from
any one sheep, let alone the variations between the sheep. Up to the 1970s, most
of the judgment regarding quality was made by appearance and touch. However,
since then there have been developments in instrumentation that have been accepted
and applied. According to Whitely [23], 90% of the variation in clean prices is
accounted for in fiber diameter and vegetable-fault level. Style categorization is an
indicator of staple length, average fiber diameter, and vegetable fault content. Scanning
systems are beginning to be used to record color, crimp, and staple length of the
greasy wool as it passes on a conveyor. This gives a glimpse of the advances which
are being made, not without opposition. Similarly difficult transitions have been
met with other fibers and in other geographic areas. However, the drop in wool
prices has reduced the resources available to promote the process of adopting
instrumented measurements.
The value of wool depends, amongst other things, on its fineness. Heavier wools
are discounted [24] and fiber fineness is an important factor in setting the price
(which is not always directly related to value). Wool prices languished through the
last decade or so, with the result that research and development was slowed. The lack
of significant growth has also been a disincentive for machinery makers to invest in
solving the difficult technological task of improving productivity. There have been
fewer developments than in the short-staple arena; there seems to be some recovery
but, as is so often the case with textiles, it is difficult to predict the outcome.
Textile products and fiber production 35
2.2.3 Bast fibers
Flax
Flax (Linum usitatissimum) is an annual, herbaceous plant grown in temperate and
subtropical areas. After flowering, the bolls or capsules contain up to ten seeds. The
fibers occur in the bark of the stem and it is the long stemmed varieties that are used
for linen. Bast stems contain bundles of fibers that act as hawsers in the fibrous layers
lying beneath the bark of dicotyledenous plants. (A dicotyledon is a plant having two
seed leaves.) They help hold the plant erect. The Soviet Union was the largest producer
before the collapse of communism but is no longer. Some satellite countries of the
former USSR, such as Slovakia, produced large quantities of flax and linen yarns,
some of which were directed to the manufacture of tarpaulins and other industrial
uses. Linen yarns have remarkable resistance to sub-zero temperatures, which cause
deterioration of properties in many synthetic fibers. Any conception that linen is
solely an apparel fiber is misguided. In fact, in common with other bast fibers, it is
beginning to find a use as a reinforcement for composite materials in automobile
manufacturing because of its strength and biodegradability. Other major producing
countries in recent times have been Poland, Germany, France, Ireland, Rumania,
Belgium, and Holland.
Strands of commercial flax fiber may consist of many individual fiber cells. The
cells vary from about 0.25 to 2.5 inches in length. They exist as thick walled, cylindrical
tubes with a diameter of about 0.0008 inch and the central lumen (the central canal
in the cell) tapers to a point towards the end of the fiber. The fibers do not have the
convolutions typical of cotton and the width of the fiber may vary several times along
its length. It is stronger than cotton but it is an inextensible fiber and the elongation
at break is only about 2%. It is 20% stronger wet than dry. Flax is still the main
vegetable fiber grown in northern Europe [25].
The plants are subject to attack by pathogenic fungi (wilt) and viruses (curly top).
Wilt-resistant varieties have been developed. Reasonable control can be exercised by
chemical treatment of the seed and the use of fungicides.
Flax is harvested when about half the seeds are ripe (yellow or brown, shiny, and
flattened) and the leaves have fallen from the lower two-thirds of the stem. Modern
practice is to use pulling machines that remove the plants bodily from the soil and
bind them into bundles which are set into ‘stooks’ or ‘shocks’ in the field for drying
or curing. The stooks are assemblies of bundles of stems arranged like an elongated
cone that promotes natural airflow through the bundles. When they have dried, the
stalks are de-seeded by threshing, combing, or beating, and the product at this stage
is referred to as ‘straw’.
Before the fiber can be used for textile products, it has to be removed from the
stems. The dried straw is ‘retted’ to break down the gums that bind the fibers together
in the bark. ‘Retting’ is a controlled rotting process which is brought about by exposure
to the weather, or soaking in ponds, sluggish streams, or vats. Bacterial action and the
physical effects of weathering or soaking cause the decomposition of the gums.
Retting is complete when the bark becomes loose so that it can be easily removed
from the woody portions of the stems. The process takes one to three weeks according
to the weather or the temperature of the water. The retted straw bundles are set up in
open shocks or ‘wigwams’ to dry. The fiber is separated from the woody material by
‘scutching’. This involves the use of fluted rolls and beating blades which break the
brittle woody parts into ‘shives’ but leave the fibers largely intact. The scutched fiber
is baled and sent to the mill.
36
Handbook of yarn production
Jute
Jute fibers are obtained from two species of Corchorus, namely C capsularis and C.
olitorius. There are also a number of jute substitutes such as Bimli (from Hibiscus
cannabinus) and China jute (from Abutilon theophrasti). Jute fabrics formed the
‘sackcloth’ of Biblical times and are now used for wrappings, bindings, etc.
Commercial jute fiber consists of overlapping cells which average 0.08 inches
long by 0.0008 inches equivalent diameter (cells are not round; the equivalent diameter
has the same cross-sectional area as the cell). The color varies from yellow to brown
with various degrees of grayness and tends towards brown when exposed to sunlight.
Like flax, the fibrous material surrounds the woody core and is embedded in the nonfibrous material under the bark. The strands nearest the bark run the full length of the
stem and other strands further from the bark become progressively shorter. The cells
are about 0.1 inches long and, although retting destroys the tissue that holds the fiber
bundles together in the natural state, it usually does not separate the cells in a given
fiber.
Cultivation requires well-drained, fertile soil and a hot, moist climate. The crop is
ready for harvesting when the flowers begin to fade. If cut too early, the fiber is weak,
and if cut too late, it is strong but coarse and lacking in luster. Like flax, the stalks are
retted to free the fibers from the natural gums that bind them. If the stems are
removed from the retting basins too soon, the fiber is difficult to remove and suffers
mechanical damage. If they are allowed to stay immersed too long, the fiber is
degraded and is weakened. The separation of the fiber is termed stripping. The
material is graded and baled before shipping to storage.
Hemp
The botanical name for hemp is Cannabis sativa. Sisal and manila hemps are hemp
substitutes. As mentioned earlier, C. sativa produces fiber, seed, and narcotics.
Cultivation is not unlike that of other bast fibers and, again, the time for harvesting
has to be judged carefully. The fibers are soft and fine if they are harvested as the
pollen begins to shed, but they are weaker than those obtained from later harvesting.
Hemp made its mark because of the strength of the fiber. The cells vary from about
0.5 to 1 inch long, and, like flax, they are thick-walled tubes, although the lumen has
blunt ends. The fibers may be up to 6 ft long and are roughly cylindrical with cracks,
swellings, and other irregularities.
The process of retting is similar to that already described for other bast materials.
The devices for separating the fiber and ligneous material are called ‘brakes’ and the
process is called breaking, but essentially the process is similar to that already described.
Ramie
Ramie comes from plants with the botanical name Boehmeria niva or B. tenacissema.
Fibers are removed by decortication, which is a process whereby the fibers are
removed from soaked stalks by scraping or beating. Gums are then removed by
soaking in caustic soda followed by neutralization in an acid bath. The fiber is then
washed and oiled. The thick-walled cells often reach 18 inches long. Normally the
fiber is rather stiff but mercerized ramie has some qualities that allow it to approach
the performance of cotton.
Textile products and fiber production 37
2.2.4 Silk
Silk fibers
Silk is different from the natural fibers previously discussed because it occurs as a
filament, and the highest quality silk is worked as filament rather than as staple fiber.
That is not to say that it cannot be chopped and mixed with other staple fibers. The
visual and tactile characteristics of silk make it very attractive in both forms. As with
all textile fibers, silk has long-chain molecules as its backbone, but attached to these
are various sorts of side chains. Crease resistance and yellowing are two problems
that have been addressed by epoxide treatment [26]. Some of the treatments are
intended to cause the fiber to behave in a more nearly elastic manner.
Silk is extruded by the silkworm into a cocoon and the silk has to be reeled from
that cocoon before it can be used. Silkworms are of the Lepidoptera family, and of the
Bombyx species, which feed only on mulberry leaves. Cultivated species are often B.
mori, but there are also other species such as B. textor and B. sinensis. Indian Tasar
silkworms are Antherea proylei and A. mylitta, which feed on leaves other than
mulberry. A. assamensis (Mugar) silkworms and others are also used for fiber harvesting.
The silk glands of the larvae produce fibroin. The freshly made fibroin is transferred
to two holding cavities for the fluid to ripen. When the caterpillar reaches maturity,
fibroin is extruded through a spinneret in common with a second secretion called
sericin. Fibroin is an amphoteric colloid protein and sericin is a natural gum. This
sericin solidifies straight away and two entering filaments of fluid are converted to a
single emerging strand that is used to cover the insect in an oviform envelope. The
filament varies from white to yellow in color. It has a high tenacity and it is capable
of 20% elastic (i.e. fully reversible) elongation, which is remarkable.
After the seracin gum has been removed, raw mulberry silk strand consists of two
smooth rod-like fibroin filaments and has a white lustrous appearance. The crosssection of each filament is roughly triangular.
Obviously, a supply of larvae is needed, but this will not be addressed here. Also,
adequate supplies of appropriate leaves are required to feed the larvae; thus the first
step is to cultivate the mulberry trees or other plants. A uniformly hot climate is
needed to hatch silk ‘seeds’ or eggs, which are usually set out in trays. Incubation is
timed to coincide with the leafing of the feed plants. After being taken from the
incubator, the trays containing the newly hatched eggs are spread with gauze on to
which chopped leaves are spread. The feeding period continues for about 40 days.
The caterpillars, to spin their cocoons, inhabit a structure of cells. This structure is
rotated whilst the cocoon is being spun. The cocoon takes about 60 hours to complete.
It is essential to keep the cocoons separate during spinning because, if they stick
together, it is almost impossible to reel the silk from the cocoon. The chrysalises are
then killed, often by steam, otherwise the pupae would damage the cocoon in emerging
therefrom. In some areas, reeling is carried out in a portion of a silk processing
establishment called a filature. The reelable cocoons are boiled in water for about
10 minutes to soften the sericin. The ends are then sought for each of several cocoons
and the group is reeled to make a skein or small bundle (called a ‘book’). Sericin has
poor solubility in the presence of tannins and this makes it difficult to degum nonmulberry silks containing tannins. Modern developments include automatic reeling
machines; also enzymatic and other forms of decomposition have been used with
some success.
The books are baled for further processing elsewhere. The count of raw silk is
expressed as denier. The inverse of 4 464 531 yd/lb is equivalent to 1 denier. The
38
Handbook of yarn production
discarded non-reelable cocoons are mechanically converted to staple and are used to
make a spun silk described as ‘schappe’. The materials from breaks in reeling are also
used for spun silk yarns. Staple silk spinning is, in some countries, a cottage industry
with spinning wheels and mules still in use. These cottage industries are often promoted
for social reasons. Throstle spinning and ring spinning are used in other areas. Matsumoto
et al. [27] describe an improved throstle in which a rolled sheet of silk fiber is
inserted into an intermittently rotating stuffer tube.
Section C
2.3
Man-made fibers (polymer extrusion and yarn production)
2.3.1 Outline of processes to produce man-made fibers
It is possible to emulate the silkworm by liquefying the polymer before forcing it
through a ‘spinneret’ to produce a number of fine streams. The streams of liquid are
then solidified to make filaments. These filaments usually have to be ‘drawn’ at a
later stage to further orient the molecular structure to give the desired physical
properties. The method of liquefying the polymer depends upon the type of polymer.
In some cases a solvent is used, and in others the polymer is melted. Polymer solutions
are solidified by removing the solvent by evaporation (dry spinning), or by coagulating
them in a liquid bath (wet spinning). Polymer ‘melts’ are solidified by cooling them
below their melting points (melt spinning). Some fiber cannot be melt spun because
the material starts to decompose before it melts, or because the melting temperature
is not within an acceptable range. For example, with some acrylic polymers the high
temperatures required for melt spinning can cause the fibers to discolor. Therefore,
either a wet spinning method is used, or a melt spinning operation is operated using
a blanket of inert gas to prevent oxidation (oxidation causes the yellowing, as well as
some other undesirable changes to the polymer).
Melt spinning is the most important of the two processes. In all cases, the liquefied
polymer has to pass through some form of pump to produce the necessary pressure
to force the material through the very fine holes in the spinneret. In commercial
production, the pressures are high. The flowing polymer has also to be filtered to
prevent lumps, such as gels and foreign bodies, from clogging the holes in the
spinneret.
Dry spinning
Dry spinning is used to produce cellulose acetate fibers from an acetone solution;
also several vinyl fibers and polyacrylonitrile fibers can be produced from solution
in other organic solvents. The first step is to produce polymer solution, which is then
filtered and pumped through the spinneret as indicated in Fig. 2.9. Choice of solvent
is governed by considerations of solvent power, boiling point, heat of evaporation,
stability, toxicity, hygroscopicity, ease of recovery, and cost. Low boiling point solvents
with high heats of evaporation may cause polymer condensation on the surface of the
filament and produce an undesirable surface.
Textile products and fiber production 39
It takes a finite time and considerable energy to remove the solvent from the
filaments in dry spinning. The process of solvent removal reduces productivity and
increases costs. This is because the mass transfer and heat transfer are far from
instantaneous. The spinning apparatus has to incorporate a long ‘chimney’ to remove
solvents (see Fig. 2.9). The volatile solvents needed for the process are nearly all
toxic and/or flammable. The vapors cannot be released into the air and they have to
be recovered. Also, solvents are expensive and it is an economic and ecological
necessity to recover them. Increases in delivery speed sometimes require disproportional
capital costs; there is also an upper limit to the production speed. Normal spinning
speeds lie in the range 800–1000 m/min. There is a limit to the length of undrawn and
unsupported filament that can be handled.
Flow through the spinneret is controlled by several factors. These include the
pressure and viscosity of the liquid. A typical solution runs at viscosities in the range
500–1000 poise; this viscosity is mainly determined by the concentration of solvent
and the temperature of the mixture. The yarn take-up speed depends on numerous
factors that include shrinkage associated with solidification.
The types of polymer and solvent affect the cross-sectional shapes of the filaments.
Rarely are the fibers round in cross-section, and changes in cross-sectional shapes
can be important in determining luster as well as other physical properties of the
fiber. When the material is chopped into staple fiber, the cross-sectional shape can
Input:
Filtered
polymer
solution
Pump
Solvent
+ gas to
recovery
system
Spinneret
Filaments
Chimney
Output: Yarn
Air or
inert gas
Godet
Ring &
traveler
Bobbin
Fig. 2.9
Dry spinning
40
Handbook of yarn production
affect cohesion between the fibers which will, in turn, affect the processability as
well as the properties of the final staple yarn.
Wet spinning
Wet spinning is a chemical precipitation process. Coagulation of the filaments involves
two-way mass transfer with the coagulating agent (e.g. acid) diffusing inwards into
the filaments and the chemical products of coagulation (e.g. salts, H2S) diffusing
outwards. It is sometimes necessary to use an intermediate process to produce a
solution. For example, in viscose rayon production, a soluble derivative (cellulose
xanthate in this case) is produced and this is dissolved in dilute dolium hydroxide to
produce the liquid suitable for extrusion. Solvent is leached out by the liquid in the
bath; the latter must be miscible with the solvent but must be a non-solvent for the
polymer. Thus, in a generalized diagram of such a process, it would be necessary to
specify the extrudate as filtered ‘polymer derivative’ or ‘polymer solution’ according
to whether an intermediate step is necessary or not, but in the case cited above, the
extrudate is a polymer derivative (Fig. 2.10).
During coagulation, several simultaneous processes occur, in different ways for
different polymer/solvent systems. Their coagulation is slow; up to 3× drawing is
possible. The more rapid the coagulation, the more inhomogeneous is the crosssection. The heat- and mass-transfers between the extrudate and the liquid of the
coagulation bath affect the temperature and solvent distributions within the fiber.
Any maldistributions make it difficult to obtain uniform properties throughout the
cross-section of the strand. The outside surface of the filaments hardens and this
tends to inhibit the mass transfer required. Furthermore, the migration of the solvent
through this hardened ‘skin’ reduces the volume of the material enclosed with the
result that the skin wrinkles. Thus, wet spun filaments usually have a convoluted
cross-section. Wet spinning is commonly used to produce viscose rayon, and
polyacrylonitrile (PAN) fibers. (It will be noted the acrylic fibers can also be
manufactured by dry spinning; see above.)
The polymer derivative usually has to be ripened because as it ages, it changes
Input:
Filtered polymer
derivative solution
Spinneret
Output:
Filaments
Coagulation bath
Pump
Input:
Filaments
Washing
Airflow
Chemical
treatment
Output:
Filament
yarn
Drying
section
Not to scale
Fig. 2.10
Ring &
traveler
Bobbin
Wet spinning
Textile products and fiber production 41
viscosity and character. Often it is necessary for there to be a storage system between
the preparation of the polymer derivative and the final extrusion in order to accommodate
the ageing process.
The viscosity of the polymer solution is an important variable. Generally, the
higher the concentration of polymer (desirable for economic reasons), the higher the
viscosity. High viscosity solutions spin well because the desirable cohesive effects of
high viscosity outweigh the undesirable effects of unavoidable surface tension which
tend to cause the liquid to degenerate into droplets. However, a high viscosity liquid
is difficult to filter and pump, and there must be a compromise. Frequently, the
solution is heated to reduce the viscosity during filtering. Cellulose fibers may be spun
at about 50°C (122°F), whereas polyacrylonitrile fibers are frequently spun at 170°C to
180°C (338–356 °F). A typical spinning speed is several hundred meters per minute.
Following the actual spinning operation, it is usually necessary to have a chemical
treatment such as neutralization of the acid from the coagulating bath, etc., followed
by washing and drying. The application of spin finish and the winding of the filament
yarn follows this operation. Wet spinning plants have environmental problems and
the counter-measures push up the costs of production.
Melt spinning
In melt spinning, the material supplied to the extruder is sometimes in a solid granular
or ‘chip’ form, especially for small operations. In this case, the chip is conveyed from
the storage silo to the hoppers of the extruders by a pneumatic transport system. From
each hopper, the polymer passes through the extruder, conveyed by an ‘auger’ or
‘screw’ (Fig. 2.11). The polymer is then melted by the heated barrel and by friction
Polymer chip input
Heating jacket
Insulation
Barrel
Auger or
screw
rotates
Filter +
pump
Quench air
Filament
output
Fig. 2.11
Simple fiber extrusion
42
Handbook of yarn production
from the screw. Heat flow is the major factor in determining the viscosity of the
liquid polymer in the working extruder. The viscosity affects the pressure generated
by the screw forcing the liquid polymer through the spinneret. Additional backpressure is generated by a filter pack. The design of the screw is a very important
feature of a modern extruder.
In other cases, the polymer can be supplied in a continuous molten form (rather
than in the intermediate chip form) direct from the polymerization reactor or an
intermediate heated storage tank. By supplying the polymer through heated pipes, it
can be maintained in the liquid form and air can be easily excluded to prevent
oxidation with its deleterious effects. The liquid polymer may be pumped through the
filter pack and spinneret by several sorts of pumps including the extruder screw.
Back-leakage, polymer overheating, and degradation by excessive working are factors
that have to be taken into account in choosing an appropriate pumping system. It
should be noted that these entire polymerization/spinning systems are very large and
the total polymer synthesis and extrusion equipment requires considerable floor space
and headroom. Usually a multistory building is required. The capital cost is very
high.
As mentioned, the rate of heat transfer from the extruder barrel into the polymer
is very important in determining the viscosity of the molten material approaching the
spinneret. This, in turn, helps to determine the flow rates and the ultimate yarn
properties. The rate of heat flow away from the extruded filaments leaving the spinneret
helps to determine the morphological structure of the yarn. Morphology relates to the
degree of crystallinity and orientation. At high speeds, the shear rate in the extrusion
zone (which is a function of the filament velocity) also affects the morphological
structure. The amount of subsequent drawing of these filaments yet further affects
the properties of the yarn. Such drawing might be carried out near the extrusion
operation, or during texturing, or both. The properties of the extruded filaments and
fibers are important but it is beyond the scope of this book; the reader is referred to
reviews by Mukhopadhyay [28] and Brunnschweiler and Hearle [29].
The mechanical process appears to be inherently simple. However, as indicated
above, there are a number of less obvious complexities involved which become very
important at the high speeds now in use (of the order of 5000 m/min). Since the
drawing speed is limited by the mechanical nature of the process, the extruder delivery
speed would become virtually fixed at a relatively low level if the filaments were
fully drawn. If part of the drawing procedure is deferred until the material is in the
texturing machine, the yarn leaving the extrusion frame is only partially oriented. The
orientation is completed by drawing in the texturing process and the use of this
strategy results in economic advantage. Draw ratio changes affect the strength of the
partially oriented yarn (POY). The strength of the POY produced at low draw ratios
is insufficient for high speed texturing and it becomes desirable to draw at the texturing
stage to increase the filament strength. Some drawing is necessary at the extrusion
stage to give sufficient strength and stability to the POY. Thus, when the POY is being
produced for draw-texturing, the texturing speeds in effect become linked to the
extrusion speeds. Draw-texturing relates to the process where drawing is carried out
at texturing. Since it is economically advantageous to do as much drawing as possible
at the texturing stage, the choice of draw ratio at spinning becomes fairly critical.
Commercial filament extrusion is more complex than indicated in Fig. 2.11. The
filters are larger and the molten polymer is distributed to groups of spinpacks fed
from a main spin distributor and pump system. The whole system is carefully crafted
Textile products and fiber production 43
to avoid stagnant flow zones, to conserve heat, and to preserve the temperature of the
melt with heat transfer fluids that are sometimes of two-phase type, which hold the
temperature at the boiling point of the fluid. An example of a two-phase system is
given in Appendix 3, but the example is not intended to imply that steam is always
used. It will be realized that it is very important to control the temperature and
viscosity of the melt because the uniformity of the yarn depends on it. For example,
with some polyesters it is necessary to hold the temperature between the limits 300
± 1°C. Mechanical working of the melt also affects the viscosity; therefore the design
of the extrusion and distribution systems is critical. The extrusion systems also have
to be made to permit the changing of filters and spinning heads with minimal interruption
to production. It has to be realized that cessation of flow causes problems and if the
polymer is allowed to solidify, this is a disaster!
The extruder
The screw and barrel of an extruder fulfill a number of functions. First, the screw acts
as a propulsion unit that transports the feed material to the spinneret. Second, it acts
as a pump in which the feed is compacted or compressed and later (after the polymer
has melted) forced through the various obstacles ahead. Third, it acts to work the melt
and to make it more homogeneous. The barrel acts as part of a heat-exchanger to
maintain the temperature of, or to melt, the moving polymer.
The molten material has to be filtered (perhaps with the generation of extra pressure)
before it passes to the spinneret. The first part of the process is part of the extruder
head and comprises the phases of propulsion, compression, heating, working, filtration,
metering, and extrusion of the polymer through the spinneret. Filtration and metering
will be discussed later. The second part follows and comprises quenching, drawing,
and winding the filaments.
Clearance between the screw and the barrel (Fig. 2.12) is of some importance. If
the clearance is too large, there is appreciable pressure loss and the molten polymer
leaks backwards down the screw. If the clearance is too small the screw may seize up.
As the barrel is heated, the bore diameter increases due to metal expansion, and as it
cools the bore becomes smaller due to contraction. Because of thermal inertia, the
barrel can cool down faster than the screw. Care has to be taken to ensure that the
contracting barrel will not grip the screw and cause a seizure. The procedures for
starting and stopping have to be carefully executed to avoid such mechanical seizures.
D
X
B
X
Screw
Clearance
A
Barrel
Polymer
Fig. 2.12
Cross-section
at X–X
Polymer flow in the extruder barrel
44
Handbook of yarn production
Of course, the extruder should be empty or the polymer in the barrel should be heated
to liquefy it before any attempt is made to start. If the polymer were to become crosslinked due to oxidation and it were no longer possible to melt it, the material would
have to be chipped out mechanically. It is therefore normal to operate extruders
continuously for 24 hrs/day, 7 days/week, to avoid these problems.
The screw rotates and the movement of the surface of the helical portion in direction D
causes the polymer to move in the direction C (Fig. 2.12). Polymer flows along the groove
in the screw, and the mass flow, Q, at any cross section such as (X-X) is given by:
Q = ρAV
[2.1]
where Q is the mass flow
ρ is the density of the polymer (defined as 1/specific volume)
A is the cross-sectional area
V is the mean velocity component in direction of flow.
As the chip is compacted, liquefied, and then pressurized, ρ changes and it is
necessary for AV to change accordingly. Consequently, the core of the metal screw is
tapered with the thick end towards the outlet. The forced flow pressurizes the fluid
polymer ready for delivery to the pump/filter system that precedes the spinneret. The
spinneret should produce one filament per hole; thus for a normal yarn, it must
contain many fine holes. The shape of the holes determines the cross-sectional shape
of the filament; however, the cross-sectional areas of the filaments vary from those
of the spinneret holes for the reasons discussed later.
Hot fluids usually circulate through channels in the barrel to provide the heat
rather than using direct heating. Heat flow into the polymer is equal to a proportion
of the heat flow from the heating medium plus local frictional heating. The absorbed
energy is carried away from the system by the polymer as a change in state from solid
to liquid and/or as a change in temperature. Heat transfer properties are also affected
by the changes in state and temperature. The frictional heating component and the
heat transferred from the heater are directly affected by the changes in the polymer
and there is a very complex interactive situation. Local pressures, specific volumes,
coefficients of friction, and viscosities of the melt vary according to the temperature
of the polymer. Since the actual extrusion through the spinnerets is highly dependent
on the viscosity of the melt, careful control is required of the variables.
The complexity of the operating conditions, plus the ambiguity in flow caused by
the multitude of parallel flow streams, leads to the possibility of uneven distribution
of the polymer flow. This phenomenon, known as ‘channeling’, can cause more
polymer to flow through some spinneret holes than others. The result is that filaments
have differences in linear density. Furthermore, the flow pattern can change continually
during operation, particularly if the polymer viscosity is incorrect. Operating under
such faulty conditions leads to quality control problems concerning ‘denier’ variations.
It is important to protect the polymer at high temperature from oxidation. Any
such oxidation causes changes in viscosity, cross-linking, and deterioration in the
final product. Consequently, antioxidants are usually included in the original polymer
chip or molten polymer supply Also, hydrolytic degradation is limited by drying the
polymer just before extrusion.
Filtering and metering
The material leaving the screw may not be perfectly homogeneous. It is particularly
important to remove any hard elements or highly viscous concentrations (i.e. gels)
Textile products and fiber production 45
from the fluid polymer stream lest they block the very fine holes in the spinneret. The
metal block in which the holes are drilled is called a die. Such blockages not only
interrupt the individual fluid streams from the affected holes but, even if the impediment
removes itself, it is unlikely that a filament will be re-established. Instead, one is
likely to get a drip (which is unoriented). Such unwanted polymer drips can coalesce
with adjacent filaments and the result is a fault that can seriously affect subsequent
operations in staple or textured yarn manufacture. For these reasons, it is normal to
filter the molten material before it reaches the spinneret. Metal webs, fabric, or
carefully graded sand is often used for this purpose, but in the latter case the body of
the filter has to be carefully constructed to preclude particles of sand from clogging
the spinneret holes. The filter introduces a high shear stress in the polymer, which
affects its viscosity and further complicates matters. In some cases, the filter assembly
is made in two parts, one of which is waiting to be used while the other one is in use.
At an appropriate time, when the pressure drop has risen or a given time of use has
expired, the second filter is substituted. This enables the first one to be cleaned at
leisure.
If the linear density of the filaments is to be maintained, it is necessary to accurately
control both the liquid flow rate and the yarn or tow take-up rate. To accurately
control the flow rate of the liquid polymer, it is necessary to use a metering device
(which is usually in the form of a pump), so that changes in viscosity and viscosity
distribution shall have little effect on the mass flow rate. The pressure upstream of the
metering pump has to be controlled so that it is not adversely affected by the metering
device. Leakages in the metering pump can also adversely affect the denier of the
filaments. Such variations are not easy to see at the extrusion stage and therefore very
careful observation and testing are required to give a high quality product. Any
variations permitted to go unchecked are likely to show up in the final product as
faults in dyeability and bulk which will create customer complaints. Diameter D1
(Fig. 2.13), and polymer density (ρ) determine the linear density of the filament.
This, in turn, is determined by the mass flow (Q) and the velocity of take-up (V). The
mass flow is the same at all cross-sections. Hence, Equation [2.1] may be applied.
It will be seen that the size of the hole in the spinneret, D1, plays no direct part in
determining the linear density of the filament. However, the shape and size of the
hole determines the flow lines in the polymer as it begins to solidify, and the shape
of the hole does affect the cross-sectional shape of the filament. Also the shape of the
hole, viscoelastic variables of the polymer, and the speed of take-up affect the ratio
D1/D3. If the polymer solidifies quickly after leaving the spinneret, these factors can
materially affect the morphological character of the filaments produced. This is
Portion of spinneret nozzle
D1
Polymer flow
Bulge
Hot pin
Heat
Fig. 2.13
Polymer bulge
D2 D3
46
Handbook of yarn production
because the crystal nucleation is affected by variations in shear stress, temperature,
and viscosity in the regions near to the spinneret holes. Also, under certain conditions,
there can be periodic changes in D2, due to vibrations within the polymer stream, and
these vibrations can lead to variations in the denier of the fibers.
Filtering does not reduce the debris generated at the exit of the extruder die and it
is possible that such debris might cause trouble in ensuing processes. Good housekeeping
at the extrusion stage is an essential ingredient of quality control.
Quenching
Liquid emerging from the spinneret has to be converted to a solid filament at a
reasonable distance from the spinneret face (i.e. the corresponding temperature has to
fall below Tm). Unless this occurs, it is very difficult to control and draw the filaments
at that stage. With slow crystallization, it may be extremely difficult to handle the
filaments once they are produced because of the lack of orientation. Hence, it is
normal to quench the newly emerging material with a low speed flow of dry air or
inert gas, usually blown perpendicular to the polymer stream. It is important to
restrict the velocity of the gas flow to prevent one molten (or semi-molten) filament
blowing into the path of an adjacent one. When such filaments touch, they will
usually cohere and produce ‘married fibers’, which can be a great nuisance. Materials
with significant numbers of married fibers or polymer drips cannot be used and are
scrapped. Equal distribution of the quench medium is also important because of the
necessity for equal cooling rates throughout the whole filament bundle. Unequal
cooling rates not only vary the morphology of filaments across the bundle but also
make some filaments more likely to break than others. This, in turn, affects the rate
of creation of undesired drips. In any case, the ill effects would show up in the final
product as changes in dye affinity.
At very high production rates, the speed of the filaments affects the quenching
rates significantly. Where it is intended to orientate the polymer during extension, the
filaments must be cooled quickly before the effects of stream orientation are dissipated.
Elongational forces acting on the viscous fluid passing into the draw-down zone,
where the semi-molten polymer changes to a solid, tend to align the molecules. If
cooled quickly, such orientation can be frozen to give a material which can be handled
and which might be, if research results can be transferred to commercial application,
suitable for use in draw-texturing.
The relative velocity of the quench air affects the Reynolds Number of the air
‘skin’ surrounding the polymer stream and this in turn affects the heat transfer or
cooling rate in the quenching phase. (The Reynolds Number is the ratio of viscous
and inertia forces; it is a dimensionless parameter useful in normalizing the mathematical
units.) The cooling rate helps determine the morphology of the POY.
Perhaps the most difficult problem concerning quenching occurs in tow production
because of the number and size of the spinnerets as well as the density in which the
filaments are packed in the extrusion zones. Also, with the large number of ends, the
chances of a break are much greater.
Filament take-off and drawing
Solidified filaments are gathered and carried from the spinning zone by devices that
grip them without squashing them. The filaments are often wrapped around rotating
cylinders or ‘godets’ and the capstan friction generated applies enough driving force
to withdraw them, draw them and transport them to the take-up system. The speed of
Textile products and fiber production 47
take-up is very high but it is very rare for the filaments to be drawn significantly at
this stage. After the draw stage, the filaments are at least partially oriented and the
delivery speed is even higher than that of the take-up. However, it is necessary to
draw the freshly extruded filaments at some stage, to orient the molecular structure
and give the desired mechanical properties. It might also be mentioned that it is
extremely important that the drawing be uniform from filament to filament and along
the length of each filament. Thus, any mechanical inaccuracies in the draw rolls
produce periodic variations in the draw. A common cause of this sort of error is due
to the uneven build-up of finish on the rolls (see next section). Unfortunately, this not
only causes variations in linear density, but also variations in dye affinity; such
variations lead to streaking and barré. The drawing of a yarn can produce filament to
filament variations in draw ratio and this can produce similar undesirable effects.
During normal drawing, a neck (Fig. 2.13) is formed, and the position of this neck
usually has to be stabilized by a hot pin or plate.
Undrawn polymers change their characteristics fairly rapidly; this is referred to as
ageing. The more the polymers have been drawn, the slower the ageing takes place,
and fully drawn filaments have very long shelf-lives. In drawing filaments, ageing of
the spun, undrawn yarn has to be controlled because it affects the natural draw ratio,
the drawing tension, and the physical characteristics of the material. The phases of
extrusion and final drawing occur at different locations when working with POY, and
the material is stored during the interim. Consequently good inventory control is
needed to keep the product within acceptable time limits between extrusion and final
drawing.
To start up a high speed drawing operation, it is necessary to use an aspirator. Such
a device sucks the yarn from the spinneret (or other source) as fast as it is produced
(it being realized that the source cannot be shut off in many cases). The ends are then
‘painted’ round the threadline and are wrapped around the take-up godet or rolls
before cutting free the material inside the aspirator – a simple operation that needs
skill to execute.
Fiber finish and treatments
Fiber finishes are necessary to lubricate the fibers or filaments and to reduce static
electrification during subsequent operations; these finishes are mostly applied by the
fiber maker. Without such ‘spin finishes’, the increase of drag due to the high coefficients
of friction might cause end-breaks or other processing difficulties. Static electricity
causes fibers to attract or repel one another, and causes some fibers to adhere to other
surfaces (such as machine parts). In either case, a high degree of static charging
causes considerable difficulty in processing. The fiber finishes can, in some cases, be
used to provide a degree of cohesion between the filaments by acting as a sort of size,
similar to that used in weaving. They also protect machine surfaces from wear and
can prevent local fusion of fibers (especially at points where the fibers or filaments
rub guides and other machine parts during high speed winding). Fiber finishes usually
comprise a base lubricant, an antistatic agent, an emulsifier or solubilizing agent, and
various special additives. The special additives include bactericides, antioxidants,
and friction modifiers. The base lubricants are usually alkyl esters of fatty acids,
hydrocarbon oils, waxes, vegetable oils, or mixtures thereof. These finishes have to
be formulated with a regard to their sorption, moisture uptake, and surface tension
characteristics, as well as to their effect on the dielectric and flow properties of the
finish. Care also has to be taken to control the volatility, smoke potential, and flash
48
Handbook of yarn production
point of the finish so that problems in subsequent processing are minimized. These
factors are particularly important in texturing, where surface temperatures of the
fiber can reach high levels. During processing, particles of finish and fiber become
detached and deposited on various machine surfaces. Two examples of the problems
caused by this may be cited. In texturing, they can form deposits on the heaters and
other working parts. In rotor spinning, deposits in the rotor can be troublesome as
discussed in Chapter 7. It is very important that the amount of finish applied to the
fiber be strictly controlled and that the nature of the debris should be such as to
minimize difficulties in the ensuing processes. Also, the finish should have no
detrimental effect on the package including its shelf-life. Sometimes the polymers
include substances such as titanium dioxide (TiO2) as brighteners or other modifiers.
Brighteners hide any tendency to yellowness in fabrics made from the fibers and
makes colors more brilliant, but sometimes the additives are abrasive. For example,
fibers containing TiO2 tend to wear guides and it becomes necessary to use ceramics
at the wear points and to avoid frictional contact as much as possible in the design of
the yarn-handling portions of the equipment.
2.3.2 Man-made staple fiber production
Tow
Fiber produced for the manufacture of man-made staple yarn is first produced as tow.
The word ‘tow’ has many meanings but in the present context it means a thick bundle
of continuous filaments. Tow has to be cut, broken, or abraded to convert it into
staple fiber. The abrasion technique is restricted to light tows and for a very few
specialty purposes; it will not be further discussed here.
Fiber makers cut tow and blend it before baling to ensure uniformity of the product.
A very large volume of fiber is cut in this way for short-staple spinners. Some longstaple is also dealt with in a similar way but some is supplied to the mill in tow form.
This tow is cut or stretch-broken in the mill. Tow intended for stretch-breaking (see
below) in the mill usually has a linear density of about half a million denier. (The
linear density of tows for use by the fiber makers for cutting into staple is many times
greater.) It is difficult to find common ground in the early processes because the
needs vary so widely, as can be seen by examining Table 2.2.
Stretch-breaking tow
Stretch-breaking is a form of drawing in which the draw ratio exceeds the breaking
elongation of the filaments, with the consequence that they break as they pass through
Table 2.2
Fiber to sliver conversion
Process
fiber
Filament tow
Bast fiber
Wool
Cotton
Man-made SS
Baled MM LS
Clean
mech
Clean
chem
X
X
X
X
X
**
Cut or break
in a mill
X
*
*
**
Open
Card
X
X
X
X
X
X
X
X
X
X
Notes: X = always, * = in some cases only, ** = rarely, SS = short staple, MM = man-made, LS = long staple.
Textile products and fiber production 49
Input
tow
Roll
stand
Roll
stand
the draw zone and create staple fibers. The most popular use of stretch-breaking is to
produce long-staple sliver from which high bulk staple yarns are made.
There are several phases in the process which will be described separately but, in
fact, all the phases are often incorporated into a single machine. The phases are: (1)
heat the unbroken filaments and cool under tension, (2) break the fibers by applying
elongation stress, perhaps accompanied by a beating action, (3) relax the fibers by
heating to create bulk in the product. There may be one or more repeat stages of phase
(2) (called re-breaking) before stage (3).
In its simplest form, the machine produces a variable fiber length and the mean
length is determined by the ratch setting (the distance between consecutive roll pairs
in a roller drawing system). To break a large bundle of strong filaments requires a
very robust set of drawing elements with strong gripping power. For this reason, the
total fiber denier must be limited simply because the load cannot exceed the gripping
power of the rolls. Also damage to the rolls has to be avoided. If load-sharing between
filaments in a disorganized bundle is poor, uneven breaking will occur. Thus, it is
desirable to have a sheet of parallel fibers entering the break zone but this is not
practicable. In the early stretch-breaking machines it was not possible to process tows
heavier than about 100 000 filaments of 1.5 dpf (denier per filament). More modern
machines can process tows of up to 500 000 filaments and the acceptable fiber
fineness has increased also. The exact specification of machine capability depends
on the fiber because the fiber properties obviously play a large part in determining
acceptable loads. The loads on the rolls are measured in tons and the machines have
to be very robust. The machines are mainly used as tow-to-top machines that produce
sliver. A ‘top’ is the name for a sliver as used in the wool processing mills.
It is usual to heat the filaments above Tg (glass transition temperature) whilst they
are under tension and allow them to cool before the tension is released (phase 1). The
heat-stretch phase of the process (Fig. 2.14(a)) reduces the breaking elongation in the
stretch-break zones and makes this part of the operation easier. After the heat-stretch
Heattreated tow
Heat flow
Sheet of tow
under tension Cold air
(a) Heat-stretch
Breaker bars
Heattreated
tow input
Roll stand
(b)
Fig. 2.14
Stretchbroken tow
output
Roll stand
Stretch-break
Stretch-breaking tow
50
Handbook of yarn production
zone come the stretch-break zones where the cooled, heat-stretched tow is broken
into staple fibers (phase 2). The locked-in extension is released when the fibers are
reheated above Tg and this process causes shrinkage (phase 3).
This is a valuable way of inducing bulk in the material. Older machines used
intersecting breaker bars to control the staple length (Fig. 2.14(b)). This practice is
declining and stretch-break/re-break systems are taking their place. The re-break
stage is merely the stretch-breaking of sliver that has already been stretch-broken
once; the second stage selectively breaks the longer fibers and reduces the variation
in length. The intersecting breaker bars have an onerous duty and wear rates are
something of a problem. Modern machines are very robust and are designed for very
high speeds. The capital cost is high but they can be cost-effective where 100% manmade fibers are to be processed and the system can be integrated into the operation
without undue disruption.
Stretch-breaking not only changes the linear density of the bundle by drawing but
it also changes the linear density of each of the filaments. The filaments are stretched
to their breaking point and this involves an elongation of the fibers. Elongation is
accompanied by a reduction in linear density of the fiber; the change in dpf (denier/
filament) can be significant. The fact that the fibers are stretched whilst heated
causes flats to form on their surfaces [30] and this gives the resulting yarns a greater
crispness in hand than otherwise would be the case.
Fiber cohesion is low in freshly broken tow and, to be able to manipulate the
material, it is necessary to improve it by fiber crimping. The usual crimper is a stuffer
box in which the sliver is fed to a heated stuffer chamber at a speed faster than the
offtake in a manner similar to that described in Chapter 4. Fibers buckle under the
compressive load and become crimped (16 to 20 crimps/inch is normal). Crimped
fibers cohere well and a sliver made of such fibers can be handled and carded
properly. Where breaker bars are used, significant amounts of fly (airborne fiber and
debris) can be generated and this fly must be taken away from the breaker zone
otherwise the product becomes contaminated, to the detriment of following knitting
operations.
One great advantage of the stretch-breaking process is that it produces high bulk
yarns [31]. Bulk is generated by the differential shrinkage of the fibers, the stage
being set for this in the heat-stretch zone. Not all fibers suffer the same tensions or
reach the same temperature with the consequence that not all of them shrink equally.
The fibers that shrink the most cause the others to become compressed along their
length; the compressed fibers buckle and the buckled fibers take up more space.
Thus, a stretch-broken sliver is naturally bulky, but the effect can be heightened by
mixing non-heat-stretched sliver with heat-stretched sliver at the drawframe and then
autoclaving (heating with steam under pressure) to produce the shrinkage required.
These bulky stretch-broken yarns are a close approximation to wool yarns and they
produce soft ‘woolly’ fabrics. The yarns are often referred to as high bulk staple
yarns. Whilst the traditional way of developing the bulk was to use an autoclave,
some modern machines have a continuous heating system attached to them, which
fulfills the same purpose. The operating temperature is about 240°F (115°C) and
steam is often used as the heating medium (see Appendix 3). Within limits, an
increase in heater temperature or draw ratio generally increases the tenacity of the
fiber, but too high a temperature leads to degradation of the polymer, which, in turn,
leads to a loss in strength. Too high a draw ratio or too low a heater temperature leads
to end-breakages (i.e. stoppages) during processing and causes increased amounts of
Textile products and fiber production 51
fly. Stretch-breaking is technically possible for tow-to-yarn systems (direct spinning)
as well as tow-to-top systems (tow-to-sliver) but the high cost of tows of suitable
quality renders the system uneconomical for direct spinning. Also there is no chance
of blending between the outputs of different machines to reduce the risk of barré. The
object lesson here is that production efficiency cannot always be reconciled with
quality of product.
Cutting tow for long staple
In cutting tow to produce long-staple fibers, a spiral cutter is commonly used, which
meshes with a smooth, hardened anvil roller as shown in Fig. 2.15(a). The tow is
spread out into a sheet of uniform thickness before passing through the cutter. The
staple length is controlled by the pitch of the cutting edges and, to a lesser extent, by
the angle at which the fibers pass through the system. Also, involuntary variations in
fiber attitude cause a spread in staple length, Fig. 2.15(c). It is possible, by altering
the angle at which the tow passes through the cutter, to slightly change the staple
length as shown in Fig. 2.15(d). Only minor changes can be made by altering this
angle and any major change requires that a different cutter be used. Any damage to
a cutting edge is likely to allow double-length fibers to be discharged and these can
cause difficulties in the following drawing and drafting processes. For this reason, the
cutting edges are not razor-like but are rectangular, and function by locally crushing
the filaments at the point of contact between the cutter and anvil roller. There is
difficulty in handling fine denier fibers if the cutter is not precisely set and in perfect
condition. Maintenance of the cutter is a vital part of the operation.
Fibers tend to be bonded along the cuts by the pressure exerted by the cutter and
this is undesirable. Therefore, the ribbon is caused to flex to create shear, which
debonds the fibers as shown in Fig. 2.15(b). Also, tow leaving the cutter has welldefined lines of weakness along each cut since there can be little fiber entanglement.
If the emerging ribbon of freshly cut tow was simply condensed, the resultant sliver
would be extremely weak. To overcome this, the sheet is sheared by a process called
‘shuffling’ as shown in Fig. 2.15(e). In the case shown, an apron is used as the bottom
element to provide a reaction to the two top rolls. The position of the cut end in the
top of the sheet is now displaced from those below as shown in Fig. 2.15(g). The
distribution of cut ends disperses the zones of weakness. Finally the sheet of cut
fibers is rolled to make a sliver as shown in Fig. 2.15(f). The elements shown in these
diagrams are often parts of a single machine so that the input is filament tow and the
output is staple fiber in sliver form.
Fiber finish and subsequently added dressings are often added in the mill to aid the
tow cutting process [32] but such additives can adversely affect the performance of
the sliver in the yarn making operation. A dressing that makes the cutting operation
easier may cause fibers to cohere in a non-uniform manner. This, in turn, may cause
unevenness in the yarns produced. The Pacific Converter type of cutting equipment,
which is the type just described, is often used to produce a sliver or top, whereas the
cutters used to produce short-staple fibers are quite different.
Cutting tow for short staple
Long-staple processing is more tolerant of multi-length fibers than is short-staple
processing. Short-staple or mid-range systems are less tolerant of fibers greater than
the ratch setting of the drafting system because they bridge the drafting zones and
either break or slip at the drafting rolls. A ratch setting is the distance between
52
Handbook of yarn production
Cutter pitch
Uncut
input
tow
Cutting edges Cut tow
Cutting
roll Pressure
bonding at
each cut
Anvil
roll
Debonded tow
Flexing breaks
Bonds along cut lines
(b) Debonding
Cut tow output
A
(c)
B
Cut lines
Cut lines
Tow cutting
Cut lines
(a)
E
C
D
F
AB, CD and EF are
fibers CD > AB > EF
Effects of varying fiber angles
L2
L1
L1 ≠ L2
(d)
Effects of varying cutter angles
Input sheet of
debonded
staple fiber
Output sheet of shuffled,
debonded staple fiber
Input sheet of shuffled,
debonded staple fiber
Roller Sliver
Vb V t
Sheet rolled into sliver
(e)
Shuffling
(f)
Rolling
Vt
After
Before
(g)
Vb
Vf
Vt > Vf > Vb
Shearing action caused by the shuffling process
Fig. 2.15
Tow-to-top conversion by cutting
adjacent sets of rolls in a drafting system (see Chapter 3). These events disturb the
flow of normal fibers. This is detrimental to the efficiency of the process and the
quality of the product. One solution is to wrap the tow around a cutter of the type
shown in Fig. 2.16, first to create pressure between the filaments and the cutting
edges and second to apply internal suction. Few over-length fibers pass into the
Textile products and fiber production 53
Air
Air
Air
Cutter
Tow input
Fig. 2.16
Fiber
+ air
Tow cutter for short staple
output and the system is suitable for producing short or mid-range staple fiber. The
fibers are baled for transmission to the mill. To permit carding, it is necessary for the
fibers to be crimped so that there is a degree of mutual cohesion, as was discussed earlier.
Tow size and quality is important; the larger the tow, the more difficult it becomes
to maintain uniform tension in processing. Lack of uniformity in thickness across the
tow sheet can cause problems and, in particular, the tendency for the sheet to fold at
the edges can lead to problems. Tow knotting is also an operational problem because
the knots have to be removed before cutting. The knot removal operation can leave
gaps in the ensuing webs which result in excessive waste.
Fiber crimping and finish
As discussed, it is normal to crimp the fibers. A ribbon of fibers can be deformed
under heat by overfeeding it into a stuffer box or by passing it through the mesh of
fine toothed gears (gear crimp) as shown in Fig. 2.17. Alternatively, tow can be
stretched hot and then cooled to lock in the extension in the manner previously
Feed rolls
Heat
Sliver
Vo output
Vi
Input
Vi > V o
(a) Stuffer box crimping
Heat
Cut tow input
Crimping gears
Crimped fiber output
(b) Gear crimping
Fig. 2.17
Fiber crimping
54
Handbook of yarn production
described. The fibers must be properly lubricated to prevent damage in the opening
and carding processes. Also, it is essential that the finish applied to the fiber should
minimize any tendency for the fibers to charge electrically due to friction suffered in
processing. Electrostatic charges interfere badly with normal processing and application
of a suitable finish in appropriate quantities is very important. As pointed out earlier,
any finishes applied to aid the conversion to sliver must not cause it to perform badly
in the yarn manufacturing operation.
General comments
Conversion of tow to sliver is a short mechanical process that can be described in
relatively few words. Nevertheless, adequate quality control involves not only the
mechanical processes but also the chemical and morphological characteristics of the
polymer and fiber finish. Brevity of explanation should not be taken to mean that any
one of the processes is unimportant.
References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
Cook, J C. Handbook of Textile Fibers, Merrow Publishing, Watford, UK, 1964.
Von Bergen, W and Krauss, W. Textile Fiber Atlas, Textile Book Publishers, New York, USA,
1949.
Terms and Definitions, 8, The Textile Institute, Manchester, UK, 1986.
Textile World, McGraw Hill, New York, annually.
Encyclopedia Brittanica, Wm Benton, London, 1964.
Tagart, W S. Cotton Spinning, Vol 1, pp 20–6, Macmillan, London, 1913.
Trolinder, N L. How to Genetically Engineer Cotton, and other papers by various authors,
Proc Beltwide Cotton Conferences, pp 165–77 National Cotton Council, USA, 1995.
Lalor, W F. Sticky Cotton Action Team Activities, 7th Annual EFS Conf, Cotton Inc, 1994.
Norman, J W, Sparks, A N and Riley, D G. Impact of Cotton Leaf Hairs and Whitefly
Populations on Yields in the Lower Rio Grande Valley, Proc Beltwide Cotton Conf, pp 102–
3, National Cotton Council, USA, 1995.
Hendrix, D L. The Relationship between Whitefly Populations, Honeydew Deposition, and
Stickiness in Cotton Lint, Proc Beltwide Cotton Conf, p 104, National Cotton Council, USA,
1995.
Sasser, P E. Automation and Validation of the Sticky Cotton Thermodetector, Proc Beltwide
Cotton Conf, p 105, National Cotton Council, USA, 1995.
Websters New Collegiate Dictionary, p 202, G & C Merriam Co, Springfield, Mass, USA,
1963.
Lalor, W F, Willcut, M H and Curley, R G. Cotton Ginners Handbook, USDA, Agricultural
Handbook No 503, p 21, Washington, DC, 20250, 1994.
Shaw, D L, Cleveland, O E and Ghetti, J L. Economic Models for Cotton Ginning, Texas Tech
University, College of Agriculture, Scientific Publication No T-1-158, Texas, USA, 1977.
Anthony, W S. Cotton Ginners Handbook, USDA, Agricultural Handbook No 503, pp 43–6,
Washington, DC, 20250, 1994.
Sutton, R M. 198 Gin Stand and Dual Roller Lint Cleaner, Proc Beltwide Cotton Conf,
National Cotton Council, USA, 1995, pp 50–1.
Mayfield, W D, Anthony, W S, Baker, R V and Hughs, S U. Cotton Ginners Handbook,
USDA, Agricultural Handbook No 503, p 237, Washington, DC, 20250, 1994.
McPhee, J R. The Mothproofing of Wool, Merrow Monographs, Merrow Publishing, Watford,
UK, 1971.
Garnsworthy, R, Mayfield, R, Gully, R, Westerman, R and Kenins, T. Proc Int Wool Text Res
Cong, III–190, Tokyo, Japan, 1985.
Wool and the Wool Trade, 2nd edn, Riverside Press, Cambridge, USA, 1955.
Carnaby, G A, Stanley-Boden, I P, Maddever, D C and Ford, A M. Mathematical Concepts
and Methods in the Industrial Utilization of the New Zealand Wool Clip, J Text Inst, 79, 1,
1988.
Textile products and fiber production 55
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
Ryder, M L. The Production and Properties of Wool and other Animal Fibers, Text Prog, 7,
3, 1975.
Whitely, K J. Quality Control in the Processing of Wool and the Performance of Wool
Textiles, J Text Inst, 79, 3, 1988.
Anderson S L. Textile Fibres: Testing and Quality Control, Manual of Textile Technology,
Textile Institute, Manchester, UK, 1983.
Anson, R. Natural Alternatives, Text Horiz, ITBD, London, Dec 1995.
Tsukada, M, Shiozaki, H and Urashima, H. Changes in the Mechanical Properties of Silk
Fabrics Modified with Epoxide and their Relation to the Fabric Construction, J Text Inst, 80,
4, 1989.
Matsumoto, Y, Harakawa, K, Toriyumi, K and Tsuchia, I. A Study of Throstle-spun Silk/
Raw-silk Core-spun Yarn, J Text Inst, 82, 4, 479, 1991.
Mukhopadhyay, S K. The Structure and Properties of Typical Melt-spun Fibres, Text Prog,
18, 4, 1989.
Brunnschweiler, D and Hearle, J W S. Tomorrow’s Ideas and Profits: Polyester, 50 years of
Achievement, The Textile Institute, Manchester, UK, 1993.
Wray, G R. New Trends in Yarn Production, Modern Yarn Production, (Ed G R Wray),
Columbine Press, Buxton, 1969.
Reid, J J. Tow-to-sliver Stretch Breaking, Modern Yarn Production, (Ed G R Wray), Columbine
Press, Buxton, 1969.
Kirk, E. Tow-to sliver Cutting Methods, Modern Yarn Production, (Ed G R Wray), Columbine
Press, Buxton, 1969.
3
Common principles
3.1
Introduction
Many principles relating to yarn production apply across the range of processes. For
example, eccentric or deformed machine elements can produce periodic errors in
filament yarns just as much as in staple yarns. An understanding of the common
principles is vital to the comprehension of the technologies involved. The spectrum
covers many technologies and yarn production systems. It includes the mechanics of
drawing, doubling, and twisting. For reasons of economy, it is useful to discuss some
of these common principles before dealing with the details of each process. Readers
unfamiliar with the industry will find reference to Appendices 1 and 2 helpful.
3.2
Twist in strands
3.2.1 Purpose of twist
In the present discussion, the words ‘twisted strand’ encompass any yarn or intermediate
product that is twisted. Staple yarns or rovings are twisted to induce lateral forces.
Friction created by these forces acts to control fiber slippage in a strand under
tension. A simple experiment can demonstrate this. Take a length of sliver and it will
be found that fibers can be withdrawn from the ends with ease. If the sliver is now
twisted tightly, it will be observed that the diameter decreases as the twist is added.
Lateral forces act to compress the strand and bring the fibers closer together. As the
fibers are pressed together, it is harder for the fibers to slip and it will be found that
it is difficult to break the twisted sliver. If the sliver is then untwisted, it becomes
weak again; failure is caused by fiber slippage.
A yarn is false twisted when a torque is applied to a running yarn, and the
consequences are highly significant. Examples include texturing, rotor spinning, ring
spinning, roving production, and yarn splicing. Distinctions between real and
false twist will be drawn in the following text and the importance of each will be
discussed.
Common principles 57
3.2.2 Twist direction
It is necessary to define the direction of twist before continuing. Referring to Fig.
3.1, the direction of twist can be determined by matching the visible surface fibers to
the center portion of the letter S or Z, whichever is appropriate. The convention is to
refer to S twist or Z twist according to the direction. It is also conventional to spin
single yarn in the Z direction but to ply in the S direction.
Fig. 3.1
Twist direction
3.2.3 Twist and flow
Twist multiple and linear density are the most important parameters in determining
the character of a twisted yarn made from a given fiber. However, in setting up a
spinning machine, the linear velocity of the strand and the rotational speed of the
twister must be set in their correct proportions to produce the required yarn. The
relative speeds are controlled by a transmission system consisting of gear trains and
belt drives. In practice, a single gear (called a twist gear) is changed to alter the
velocity ratio, which gives the required rate of advancement of the strand for the
given spindle speed. The rest of the transmission is typified by a twist constant for the
machine. Twist density is often measured in twists/inch (tpi), which is calculated
from the ratio of the twist constant and the number of teeth in the twist gear. Twist
density can also be measured in the metric system. Some simple examples of such
calculations are given in Appendix 2.
3.2.4 Effect of twist on a staple yarn
Consider a simplified model of a staple yarn where a number of fibers at a given
radius exist in a roughly helical configuration. Each fiber is under tension and a
series of resultant forces acts towards the center of the fiber bundle (see Appendix 5).
Taking all the fibers at a given radius, there is a sort of tube of fibers, all of which
press inwards on the bundle of other fibers inside the tube. If a sufficiently high
pressure is maintained on the inner fibers, there can be little fiber slippage and the
whole structure becomes capable of bearing load. This requires that at least the outer
shell of fibers should be kept under tension. One way to do this might be to tuck in
the ends of each of the outer fibers rather in the manner of cord ends tucked in when
whipping the end of a fishing rod. Although this is impracticable for yarn production,
58
Handbook of yarn production
Yarn tenacity
it serves to show a principle. In fact, the actual process of spinning causes portions
of the outer fibers to be entrapped in the inner structure. Fiber migration is the name
for this phenomenon and it causes the whole of the structure to interlock. In an ideal
yarn, if the fibers were totally unbreakable, the strength of the yarn would increase
with twist in the manner indicated by the cohesion curve in Fig. 3.2. Yarn failure
occurs because the fibers slip over one another. On the other hand, in a yarn where
the fibers cannot slip but must break, the strength would decline with twist. This is
because of the reduced components of fiber tension resisting breakage as the twist
angle increases1; it is shown as the obliquity curve in Fig. 3.2. ‘Strength’ is usually
quoted in normalized units; it is then referred to as ‘tenacity’. Tenacity is the quotient
of force and linear density. The linear density is often quoted in tex and the units
commonly used for tenacity are mN/tex. It has the virtue that the value is similar for
all staple yarns made from a given fiber and twisted to the optimum degree, irrespective
of linear density. It is, therefore, a useful comparative measure of the strength of the
yarn. As will be discussed later, twist is usually quoted for this purpose as twist
multiple (TM). The manner of calculating TM is dealt with in Appendix 1. This is
also a normalized value and the yarn characteristics vary little for a given fiber
length. Thus, diagrams similar to Fig. 3.2 serve as models for all staple yarns made
from fibers of a given length and strength.
Real yarns are made from fibers that can be broken and which do slip. The
strength of a twisted bundle varies, as shown in the lower curve of Fig. 3.2. It will be
seen that twist weakens a fiber bundle by making the fibers oblique with respect to
the yarn axis; too much twist seriously weakens the yarn. This eventually overwhelms
the increased cohesion at twist levels above the optimum. The so-called obliquity
curve refers to fibers oriented at an oblique angle. The yarn strength has an optimum
value that is less than the sum of strengths of all the fibers in a cross-section. Above
the optimum twist, the yarn fails by fiber breakage and a distinctive snap can be
heard when the yarn breaks. If there is appreciable fiber slippage during the breakage
because the twist is well below optimum, no snap can be heard. Sometimes, where
strength is unimportant, yarns are produced at less than the optimum twist for economic
Obliquity
curve
Cohesion
curve
Actual result
Optimum twist
Twist multiple
Fig. 3.2 Effect of twist on yarn strength
1 Simple trigonometry shows that the component of tension contributing to strength = T cos β,
which indicates that the helix angle of the fiber (β) is very important in determining the strength
of the yarn. When β is very small, as in the case of some filament yarns, an anomaly arises
because of maldistribution in loads between fibers. The yarn tenacity at zero twist may be
slightly less than that achieved when the yarn has producer twist.
Common principles 59
reasons. Sometimes the twist used is below optimum to give a soft hand to the
product.
The cohesion curve can be changed by altering the staple length, l, of the fiber or
by altering the effective coefficient of friction, µ. The latter is altered by varying the
fiber lubricant (i.e. fiber finish) or the crimp level of the fiber. If l or µ is increased,
the cohesion curve moves from curve D to curve C along path x in Fig. 3.3. The actual
tenacity curve also alters to reflect these changes. Providing the obliquity curve
remains the same, the optimum moves from point B towards point A. It will be seen
that the optimum twist level is reduced and the maximum tenacity is improved by
increasing the staple length or the interfiber friction. This explains why a premium is
placed on the longer staple fibers and why short fibers are often removed from the
material to be spun. As will be discussed in Chapter 8, there is a limit to how long the
fibers can be before there are processing difficulties. However, there is another limit;
fibers beyond a certain length add very little to the resistance to fiber slippage.
For indirect count systems twist density = TM √N and in the direct systems twist
density = TM/√n both measured in twist/unit length.2 The symbols ‘N’and ‘n’ stand
for indirect yarn count and the direct yarn count (or linear density) respectively as
defined in Appendix 1. Some people use alpha to describe the metric version of direct
twist multiplier.
Since TM describes the nature of the yarn, it does not vary greatly within a units
system (see Table 3.1). It will be realized that the twist density required (tpi) is
strongly dependent on yarn count. Thus, to set up a spinning machine to make a
certain class of staple yarn, it is necessary to know the specified TM and count of the
yarn before the twist density can be calculated. The need for twist rises with count;
this is why fine yarns are more expensive than coarse ones (the adage ‘twist costs
money’ comes to mind).
O
Yarn tenacity
C
A
D
B
X
Twist multiple
NB O = Obliquity curve
C = Cohesion curve 1
A = Actual curve 1
X = Path of optimum tenacity
Fig. 3.3
D = Cohesion curve 2
B = Actual curve 2
Optimum twist
2 ASTM D861 uses twist density in t /cm rather than t/m.
60
Handbook of yarn production
Table 3.1
Typical twist multiples
System
Length
Use
Ne cotton
count
Nm metric
count
Nw worsted
count
Cotton
Short
Warp
Filling
Hosiery
4.0–5.0
3.2–3.8
–
120–150
110–115
–
–
–
–
Cotton
Long
Warp
Filling
Hosiery
3.4–3.8
2.5–3.0
2.2–2.6
100–115
75–90
65–80
–
–
–
Wool
Long
Warp
Filling
Hosiery
–
–
–
65–75
55–65
45–55
1.8–2.0
1.5–1.8
1.4–1.5
3.2.5 Ply twist
Where durable and pliable yarns are required, it is the practice to twist several yarns
together and this is called plying. A singles yarn might, when relaxed, tend to take up
a shape sketched in Fig. 3.4(a). It is then termed twist lively. It is usual to ply in the
direction opposite to that in which the component strands were originally twisted so
that the resulting plied yarn is no longer twist lively. Fibers on the outside of the ply
normally appear to be roughly parallel to the axis of the yarn. Such a yarn produces
a result similar to that in Fig. 3.4(b). If the amount of ply twist used is just sufficient
to remove any residual torque (i.e. the plied yarn is non-twist lively, or ‘dead’), the
yarn is said to be balanced. Such balanced plied yarns are useful in reducing difficulties
in handling the yarn during any post-spinning processes, as well as in lessening
distortion of knit fabrics.
Twist structure is normally described in shorthand fashion. A singles yarn of count
Ne = 20 is usually written as 20/1 (or 20s). When two such yarns are twisted together
to make a plied yarn, the equivalent count3 is 10equ. The ply yarn is described as a
20/2 (but in some areas it is described as 10/2, which is meant to indicate 10equ/2). If
four 20/1 yarns are plied, the result is a 20/4 yarn that has an equivalent count of 5equ.
The designation ‘equ’ stands for equivalent and it is usually omitted, which is confusing.
It is useful to check the context before working with yarn numbers for plied or cabled
yarns. With worsted yarns, the order of the numbers is usually reversed; if two yarns
(a) Twist lively
Fig. 3.4
(b) Non-twist lively
Twist liveliness
3 See Appendix 1 for calculations. There are alternative designations to indicate that the number
refers to equivalent count.
Common principles 61
of Nw = 40 are plied together the result is designated 2/20equ or 2/40. The ply twist
multiple is calculated on the equivalent count.
When plied yarns are twisted together to produce a complex structure, this is
referred to as a cable. Such cabling has a structure that is much more flexible than a
simpler one. An example will illustrate how the twist structure is designated: if six
20/2 yarns are twisted together, the result is a cable, which is sometimes described as
20/2/6; the equivalent count is 10/6 = 1.66s. In some areas, the numbers are written
in a different order. With cables, there could be an ambiguity and care should be taken
to check the context.
3.2.6 Twisted filament yarns
It is unnecessary to twist continuous filament yarns to impart strength; nevertheless,
some small amount of twist is inserted merely to control the fibers. An untwisted
bundle of filaments is difficult to handle because odd filaments and loops project
from the surface of the bundle. These tend to catch up in guides, tangle with adjacent
yarns, or otherwise cause difficulty. Some man-made fibers tend to balloon out quite
severely because they accumulate electrical charge. Filaments or loops protruding
from the yarn are often called wild filaments. Even a low level of twist in the yarns
helps to reduce the number of these wild filaments; twist inserted for this purpose is
called producer twist.
Filament yarns are sometimes twisted to a fairly high level to break up the luster
of the yarn or to impart some other attribute to the yarn for effect purposes. However,
high twist levels reduce the tenacity of the yarn and make the yarn leaner (i.e. have
a smaller diameter).
Another use of twist in filament yarns is to create texture. A false twisted yarn will
coil or snarl if it is subject to the correct sequence of twist, set, and untwist. If
properly relaxed, these textured yarns become bulky and have many desirable features.
A major advance was made when it was realized that the process of false twisting
provided the opportunity to carry out such a sequence in a continuous manner. To
understand how that works, it is necessary to be knowledgeable about false twist.
3.3
Twist insertion
3.3.1 Real and false twist
So many practical cases involve false twist that it is thought desirable to discuss it in
its own right. It is necessary first to discriminate between real and false twist. First,
let the word ‘strand’ be used to widen discussion. It is used here to include not only
yarns, but also rovings and possibly other forms of intermediate strands. We return
now to the subject of false and real twist. Real twist is created when a ‘crankarm’ of
a strand is rotated about an axis to insert twist and the material delivered to the takeup package retains all the twist, as shown in Fig. 3.5(a). The theoretical twist remains
constant from the point A until the yarn is wound onto the package at a level τ =
U/V, where τ = strand twist in tpi, V = linear speed in inches/min, and U = rotational
speed in r/min.4 Some of the practicable means of achieving this are described later.
4 Fig. 3.5(a) does not show a balloon or yarn package so that the diagram may also cover two-forone twisting.
62
Handbook of yarn production
A
A
Vτ r /m
Flow
velocity
along the
strand
axis = V
False
twister
B
Twist
flow
U r/m
C
C
Exit strand is
twisted
Torque produces
real twist
(a)
Fig. 3.5
Exit strand is
not twisted
False twist
level = τ tpi
(b)
Real and false twist
A typical example of this mode is in ring spinning, where the point C is in the upper
reaches of a yarn balloon. The yarn package (not shown) rotates at a speed similar to
that of the yarn in the balloon. In consequence, real twist extends over all the length
AC.
False twist is created when a strand flows through the torque producing means. No
twist exists in the strand delivered and the false twist is locked in the system above
the twister while the strand continues to run. Systems involving both forms of twist
are possible, in which case there is twist in the strand delivered but it is different from
the value upstream.
3.3.2 Mechanics of false twist
In the simplest form of false twisting, neither the supply nor take-up packages are
rotated to insert twist. Thus, the net twist in the delivered strand is zero and twist is
trapped upstream. To explain how false twisting works, consider Fig. 3.5(b). A is just
below the nip of the feed rollers, B is the point at which twist is applied, and C is the
nip of the take-up. Twist carried downstream from B by the yarn is –Vτ twist/min, the
amount projected by the twister is +U twist/min, and the twist delivered is the sum of
them. The difference in sign is because the false twister generates S twist on one side
and Z twist on the other. Twist flow input to zone AB is U twist/min and the loss is
Vτ twist/min.
Common principles 63
Thus the total twist in zone AB of length L is τL + (U – Vτ)t, where t = time.
(U – Vτ) must be zero, otherwise the twist level would change and the system would
not be stable. The input to zone BC is –(U – Vτ) twist/min, which results in no twist
in zone BC. Thus, the running false twisted yarn behaves as if twist was inserted at
A and removed at B.
3.3.3 False twist insertion
Only examples of false twisting may be discussed at this stage. Some are by design
and some are involuntary. Some uses are aimed at texturing the yarn and some are
aimed at temporarily strengthening the strand during processing. For brevity, only
one example will be given of each.
In texturing, twist is deliberately inserted by stacks of discs. The shape of the
filaments and the structure of the twisted yarn are frozen and then the twist is
removed, leaving the fibers in a stressed condition. The fibers are separated and the
stress is relaxed, which causes the filaments to texture themselves to create a bulky
or stretchy structure.
In roving, rubber grommets are used at the top of the flyer to provide a combination
of false and real twist in the weak strand leaving the drafting system and entering the
twister. The rubber surface grips the roving, which is in contact with the inner top
surface of the rotating grommet and they move at different speeds. The shear from
this pumps twist into the section between the drafting system and the grommet; this
is in addition to the twist generated by the rotation of the flyer (discussed in Chapter
6). In this case, the outgoing twist is similar to that determined by the rotational and
linear speeds; however, this does not imply that the structure of the strand is unaffected
by the false twist. Often of major importance is that the twist in a vulnerable zone is
enhanced and the strand is temporarily strengthened. The effect reduces end-breakages
with beneficial economic consequences.
3.3.4 Real twist insertion
Ancient systems of twist insertion were discontinuous; the yarn did not flow through
the machine in the steady and continuous manner employed by modern systems.
However interesting these ancient systems might be, space limitations preclude any
discussion of them. Conventional systems have endured for some 200 years and they
are very well developed as commercial processes. They appear in several forms.
Commercial twisting is always carried out as the textile material passes from the
supply to some sort of twisting. Where twisting is the sole objective of the operation,
either up-twisting or down-twisting may be used. It is usual to use down-twisting
when other operational phases are involved in the process (such as drafting).
Two of the most common forms of down-twisting are flyer spinning and ring
spinning. As the first example (Fig. 3.6), consider the production of singles staple
yarn. A stream of fibers is supplied from a drafting system, then twisted, and the
resulting yarn is wound onto a bobbin situated inside the yarn balloon. The second
example is of up-twisting, where yarn is withdrawn from a package before further
processing (Fig. 3.7).
Alternatives capable of economic exploitation have been sought and some of these
newer developments are discussed later. There may or may not be another process
involved. One of these involves two-for-one twisting as shown in Fig, 3.8 (discussed
64
Handbook of yarn production
Yarn
Balloon
Bobbin
NB Yarn winds on
Fig. 3.6
Down-twist (down-twist is used to wind yarn on to a package)
Yarn
Pigtail guide
Yarn rotates at ωy
Lightweight
flyer rotates
at ω f
Bobbin
rotates at ωb
Wind-off speed = ω f – ω b rb
Fig. 3.7
Up-twist (up-twist is used to wind yarn from a package)
Common principles 65
in Section 3.3.5); one yarn to be twisted is taken from a package situated inside the
two concentric yarn balloons and is wound onto another one after twist has been
inserted. An extension to this is found in plying (doubling), where the input consists
of two or more yarns and the output is a single composite strand in which the input
yarns are twisted together. The twisted yarn is usually wound onto a cheese or cone
and this gives an economic advantage.
Now let us consider the twisting as an abstract idea somewhat remote from the
supply and take-up systems. With conventional technology it is necessary to rotate a
package about the axis of the yarn to insert the twist. Depending on whether the
upstream or downstream package is rotated, we have either up-twisting or downtwisting. In such cases, it is normal to have a common axis for both the rotating
package and the yarn balloon, as shown in Figures 3.6 and 3.7. By having a common
axis, it is possible to wind yarn onto (or to unwind yarn from) the rotating package
with the same motion used to put in the twist. This makes a very neat design of
machine but it requires that the yarn package inside the balloon be small in diameter
and limited in height. Also, the yarn has to be laid in organized layers as nearly
parallel as possible to its neighbors; this is important if the process is to run smoothly
at high speed.
Down-twisting is the most commonly used of the choices enumerated. The restrictions
imposed by the balloon are substantial. (The details appear in Appendix 9 because the
analysis is rather complicated.) To make the system work, there must be not only a
twisting system but also a system to control the build of the package. Control is often
applied by the motion of the ring-rail as discussed in Chapter 7. The element that is
changed to alter the build is called a lay gear.
Up-twisting combines twisting with winding and results in a change of package
shape. In up-twisting, the package from which the yarn is drawn is rotated to insert
the twist. The receiving package is rotated only to wind the strand; thus there is a
beneficial separation of winding and twisting. Control of the yarn leaving the package
is necessary, otherwise the rate of unwinding might vary with the consequence that
the yarn tension would vary. Ideally, a constant tension is required so that a stable and
uniform yarn cheese or yarn cone is built. Physical control of the yarn balloon can be
achieved by use of a ring and traveler or by a tiny wire flyer such as that sketched in
Fig. 3.7. The flyer or traveler rotates at a speed slightly different from that of the
bobbin. The wind-off speed = ± (ωf – ωb) rb inches/min, where ωf = flyer speed in
rads/min, ωb = bobbin speed in rads/min and rb = (1/2) × diameter of bobbin in
inches. The idea is somewhat similar to that used in down-twisting.
3.3.5 Two-for-one twisting
The foregoing cases relied on at least one package being rotated to put in twist.
However, there is a possibility that requires no package rotation to insert twist.
Consideration of the case sketched in Fig. 3.8 will show how this may be done. If the
yarn is doubled back on itself to make a loop, which is rotated, then one turn of A
inserts one turn of twist in each of legs B and C. Furthermore, the direction of twist
is the same in each with the result that the twists add together. With this arrangement
there is no need to rotate either package to twist the strand; one revolution of the
spindle puts in two turns of twist. Such machines are known as ‘two-for-one twisters’.
The problem is that the large package has to be held inside the yarn balloon.
The concept first started to be used for tire cords in the 1930s, but it was a further
66
Handbook of yarn production
Yarn rotates about
XX to form a balloon
enclosing the supply
package
C
X
Yarn from
supply
package
Yarn to
take-up
package
B
A
X
Fig. 3.8
Two-for-one twisting
20 years before it came into commercial use for staple yarns. Penetration of the shortstaple market took another 20 years. These machines are now used for carpet, industrial,
and other yarns, as well as for sewing threads. The attraction is that, with a two-forone twister it is not necessary to rotate a large, heavy package to insert twist; consequently
high twisting speeds can be used. However, the problem of suspending a non-rotating
yarn supply package inside the yarn balloon leads to some mechanical design difficulties.
It also leads to a certain awkwardness in piecing up because of the relatively complex
threadline path. It is normal to use a compressed air system to blow the new end
through the fairly complicated passageway. A tension control disk, coaxial with the
package(s), gives stability to the large balloon (see Chapter 9).
It could be inferred that high speed yarn balloons of large diameter are needed.
However, high yarn tension results from various combinations of large balloon diameter
and high speed. The result is that two-for-one twisters are normally used for twisting
relatively strong strands. Some machines are used for plying, in which case two yarn
cheeses may be mounted coaxially inside the balloon. The reader is referred to a
review by Lorenz [1]
3.3.6 Wrap spinning
The direct cabling machine has one sort of wrapping spindle arrangement; other sorts
are based on the ring frame in which a hollow spindle is used.
In wrap spinning, one or more small strands are wrapped around a core yarn.
Basically, one or more yarns are wrapped around a core yarn so that the fibers in the
outer sheath differ from those in the core. The wrapping may be a strong filament or
yarn to enhance the yarn strength or the wrapping might create a texturing effect. The
core often has inferior properties and the system offers financial incentives as well as
possibilities of enhancement of the technical or visual properties of the yarn. One
Common principles 67
sort of wrapping machine has an arrangement based on the two-for-one principle.
Another is based on a ring frame in which a hollow spindle is used.
3.4
Confined and non-confined systems
The means of twisting so far considered have required that a package be confined
within a yarn balloon. There are several systems that are not so restricted and these
will be discussed next.
3.4.1 Open-end spinning
If the number of fibers in the flow cross-section is sufficiently reduced as they flow
from one package to the other, it is possible to create a so-called open-end. Such an
open-end may be rotated about the axis of the yarn to put real twist in the yarn
without great interference from the incoming fiber. It is no longer necessary to rotate
either package to twist the yarn. This is known as open-end (OE) spinning.
The process involves the separation of fibers by a severe drafting action, followed
by re-condensation. This is discussed further in Chapter 7. In OE spinning, the staple
fiber flow is separated by drafting so that individual fibers (or small clumps of fibers)
are added to the ‘open-end’ of the forming yarn (shown diagrammatically in Fig.
3.9). A rotor is normally used to collect these fibers and support the open-end. This
is discussed in Chapter 7 and Appendix 10. Yarn can be spun providing there is a
steady flow of clean fibers into the moving rotor and the yarn is continuously removed.
There is no significant yarn balloon; there is neither ring nor traveler. The result is
that the package size is limited only by the ability to wind the package. Also, the
speed is not limited by a traveler. Consequently, an OE spinning machine is capable
of high production rates. In fact, an OE ‘spindle’ is capable of producing up to ten
times as much yarn per hour as a ring spindle and, as a result, this process has become
very important.
3.4.2 Alternating twist systems
It is possible to insert twist into one or more parallel strands by using a pair of plates
Wind
Open-end
Twist
B
A
Draft
Fiber transport
Feed
Discrete fibers are detached from the feed at A by
the drafting system, transported and then added to
the open-end of the yarn being made at B.
Fig. 3.9
Open-end spinning
68
Handbook of yarn production
pressed into contact with a strand, as sketched in Fig. 3.10. Alternatively, a pair of
rolls can be made to move parallel to their axes to produce a similar effect. If the
process is to be continuous, the plates or rolls have to oscillate in the direction of the
arrows.
In the case of woolen yarns, the strand is called a roping. The twist cancels after
passing through the reciprocating rolls but sufficient cohesion between the fibers is
generated by the process to give the strand enough strength to carry it to the next
stage of processing.
In the case of self-twist yarns, the rolls are called ‘shuffling rolls’. They rotate to
deliver yarn and at the same time they oscillate parallel to their axes to produce two
(or more) strands, each of which now contains alternating twist. The component
yarns are in close proximity to one another. A length of newly twisted yarn has an
unbalanced torque (i.e. it is twist lively). If two such strands of the same twist
direction are brought together in close contact along their length and are given freedom
to rotate about their common axis, they will ply themselves in the opposite direction
to relieve the unbalanced torques. The resultant ply tends to be balanced.
The shuffling roll in the self-twist machine puts in twist that alternates from Z
through zero to S, back through zero to Z, and so on. Two such strands placed
together so that the Z twist is opposite Z twist and S twist is opposite S twist, ply
themselves to give S ply through zero ply to Z ply and so on. The result is a ply yarn
in which the direction of the ply alternates. Again, twisting and winding are separated
with the result that large packages of unbroken yarn can be made. A much higher
processing speed can be attained than with other devices, because there is no conventional
spindle or rotor.
3.5
Twist evenness
3.5.1 Torsional stiffness
An uneven yarn has a varying torsional stiffness; if a torque is applied to a length of
such yarn, some portions will become more twisted than others. Torsional stiffness of
a yarn is dependent on the yarn diameter, the disposition of the fibers in the crosssection, and the torsional stiffness of the fibers. (Torsional stiffness of the fibers
depends on their cross-sectional shape and their modulus of elasticity.) The torque
might remain constant along the length but there can be rotation of one segment with
respect to a neighboring one, which is controlled by the torsional stiffness; the result
is shown in Fig. 3.11. This phenomenon is known as twist migration; the perception
Yarn
Plates oscillate in opposition to one
another to produce alternating twist
in the moving yarn.
Fig. 3.10
Alternating twist
Common principles 69
Fig. 3.11
Twist distribution in an uneven yarn
is of twist running to the thin spots (which is largely true, but changes in torsional
stiffness can produce similar effects).
3.5.2 Variation in behavior of twisted strand
Clearly, an uneven twisted strand, such as roving, has varying diameters and hardness;
consequently it has variable behavior as it goes through a drafting system, which
results in variable twist densities. This is a particular problem with roving. where
highly twisted compact portions of the strand (called ‘hard ends’) enter the drafting
system of a ring spinning machine and cause slubs and end-breaks because of the
conditions just described. Two means of controlling this problem in ring spinning are
(a) to obtain as even an input strand as possible, and (b) to use as low a roving twist
as possible. Too low a twist will not run on the spinning system concerned. Also, in
texturing, migration of twist can disrupt the structure and create faults.
3.6
Tension control
3.6.1 Axial movement
Moving strands are nearly always under tension and the tension needs to be controlled.
If the strand is flowing along its axis, there are two main simple alternatives for the
creation of extra tension for control purposes. One is to use an additive tensioner
(Fig. 3.12(a)), and the other to use a capstan tensioner (Fig. 3.12(b)). A device can
use both methods. In the additive system, the drag from the tensioner is simply added
to the existing upstream tension. In the capstan system, the wrapping of the yarn over
F
To
Ti
To = Ti + µF
F
(a) Additive tensioner
Ti
p
To ≈ Ti eµθ
θ
To
θ is measured
in radians
(b) Capstan tensioner
Fig. 3.12
Tension controllers
70
Handbook of yarn production
the segment of subtended angle, θ, produces a transverse pressure, p, and creates a
frictional restraint to flow. The appropriate equations are given in the diagram.
Note: T = tension, µ = coefficient of friction, θ is the angle subtended by the yarn,
and e = 2.718 (the base of naperian logarithms), and the subscripts i and o refer to the
input and output respectively. It must be pointed out that the coefficient of friction,
µ, is not a fixed value and varies according to the velocity of sliding. Various factors
have an effect and these include: (a) temperature of the surfaces, (b) moisture content
of the fibers, (c) additives applied to the fibers, (d) hairiness of the yarns, (e) condition
of machine surfaces, and (f) presence of contamination.
3.6.2 Orbital movement
If a yarn orbits an axis it is subjected to centrifugal force that makes it ‘balloon’. This
is a complex subject that is addressed in Appendix 9. The tensions created are lessened
by coaxial control rings, which confine the balloon. In some applications, such as
rotor spinning, the rotating element supports some of the fibrous assembly.
3.7
Drawing
3.7.1 Terminology
Historically, the term ‘drawing’ was used in connection with the drawframe in staple
spinning. ‘Drafting’ was used regarding roller drafting systems in roving and ring
spinning. Upon the appearance of man-made fibers, the term ‘drawing’ was also used
to describe the elongational process to improve the molecular orientation of the
filaments. Custom still insists on the use of the historically founded words but in
essence there is little fundamental difference between drafting and drawing.
Linear density is defined as mass per unit length of a strand or along the flow path
of a stream of fibers.
3.7.2 Purposes of drafting or drawing
Drafting occurs when a stream of fibers passes through an acceleration zone5. The
place where the acceleration occurs is called a ‘draft zone’ and it is necessary to
control the fiber flowing through it. The solutions to the problem of fiber control are
diverse and only a few examples can be given to illustrate the importance of mass
flow control by passive devices.
There are two major reasons for drafting or drawing, which are (a) to better orient
the molecules or fibers in the strand, and (b) to change the cross-sectional area of the
strand6. In the drawing of polymers, one very important objective is to orient the
long-chain molecules to give the filament better properties. In staple processing, an
important objective is to orient the fibers within the strand by causing them to slide
over one another to give the strand better properties. It should be noted that improved
orientation can only be achieved by drafting the strand to give a smaller output crosssection.
5 Conversely, when a stream of fibers passes through a deceleration zone, condensation occurs.
6 In staple spinning, drawing is sometimes considered to include doubling.
Common principles 71
There are cases that are not always regarded as drawing but which really are. For
example, in extrusion, the linear density of the molten polymer approaching the
spinneret is higher than the sum of the linear densities of the output filaments even
before conventional drawing. The speed of the output material is faster than that of
the input. While an extruder is not regarded as a drawing machine, it always is.
3.7.3 Control of flowing material
Both polymer and staple drawing and drafting have instabilities in flow. Control is
exercised by imposing restraints on the systems. With polymer in the solid state,
control is exercised by hot pins or the like. Heat flow from the control surface permits
control of the local visco-elastic constants of the polymer in such a way as to promote
stability. In the case of staple processing, the variable frictional forces between the
flowing fibers are a strong factor in producing the instability, which reduces their
value in both yarn and fabric forms. These instabilities produce quasi-random errors
in the product. The addition of an external retarding force to the flowing fiber reduces
the instability.
3.7.4 Principle of drafting or drawing
Consider a sample of the input material before and after discontinuous drafting or
drawing. If there were no losses in the process, the mass of the input sample would
be the same as it is after drawing. Let ρ be the packing density (not to be confused
with linear density), a the cross-sectional area, l the sample length, ρi a i l i be the mass
in the input sample, and ρoaolo be the mass after drafting. It follows that:
ρiaili ≈ ρoaolo
and if the packing density is constant,
aili ≈ aolo
[3.1]
For the purely theoretical case, the change in cross-sectional area is inversely proportional
to the change in length. This is discontinuous drafting. However, in production, the
process of elongation takes place continuously with the input and output mass flows
nominally constant. Thus, the formula of Equation [3.1] can be restated to say that the
cross-sectional area is inversely proportional to the speed ratio. In practice, this is
modified by changes in the packing density and small losses have to be taken into
account, but it forms the basis of all drafting and drawing.
3.7.5 Drawing in staple fiber processing
In staple spinning, the material flows through the drafting or drawing zones of the
equipment. (The term ‘drawing’ is often used to describe the particular overall process
but it is common to refer to the components that carry it out with the adjective
‘drafting’. Thus we speak of drafting rolls and draft in a drawframe which seems odd,
but that is the common usage.)
Fibers are accelerated as they pass through each zone. Also fibers can, and do,
migrate with respect to one another along the direction of flow. Conventional theory
has been mainly restricted to roller drafting, in which there are fiber acceleration
zones within the spaces between two consecutive sets of rollers. (A similar idea
72
Handbook of yarn production
applies to filament drawing but godets are used rather than rollers. Godets are cylinders
about which a yarn is wrapped to grip the yarn for the purpose of elongating it.)
However, fundamentals merely require that the exit material moves at a greater velocity
than the entry material. The theory in Appendix 8 seeks to include the case where
fibers are drafted by toothed rolls.
3.7.6 Cumulative draft
It is not possible to achieve sufficient drafting or drawing in one step; consequently
most systems use multiple, consecutive draft or draw zones (Fig. 3.13). As shown in
Appendix 1:
∆ = ∆1 × ∆2
where
[3.2]
∆ = total draft ratio
∆1 = draft ratio in stage 1
∆2 = draft ratio in stage 2.
NB The term draft ratio is technically correct but it is frequently shortened to ‘draft’.
In staple spinning, there are usually two zones. The first (or break-draft zone) has
the function of breaking frictional bonds which form in roving (or other strands) due
to (a) setting, (b) fiber migration, (c) fiber crimp, or any combination thereof. Newly
drafted material is easier to draft immediately after such an operation even if the
break draft is small because the crimp gets set over time, and the fibers no longer
slide over one another as smoothly as freshly drafted material. The break draft varies
according to the type of fiber and the linear density of the strand; it usually varies
between 1.1 and 1.4. Overall draft is the product of the break and main drafts and it
varies from about 6 to 30 according to the machine concerned. In polymer drawing,
there is often more than one stage of drawing (perhaps using different machines) to
complete the total process and the mathematical treatment is the same as for drafting
in a staple process. However, one would use the term drawing rather than drafting.
Nevertheless, for simplicity the explanation will be expressed in terms of draft.
Normally, it is arranged that there is little change in fiber characteristics, to prevent
the need to change the draft program and hence unnecessarily escalate costs.
For more than one stage, all the drafts are multiplied together to give the overall
draft. In staple spinning, the process starts with a bale laydown that might be regarded
as an extremely thick strand (a linear density of perhaps a billion (109) tex). The yarn
leaving the mill may have a linear density of less than 102 tex. (1 tex = 1 g/km or
1 mg/m as discussed in Appendix 1.) The mill can be regarded as a gigantic complex
Input
Drafting rolls
Output
Thick
Slow
Thin
Fast
Break draft = ∆i Main draft = ∆o
Fig. 3.13
Draft distribution
Common principles 73
drafting system and it is clear that a drastic amount of drafting is needed over all the
various machines in the production line. Although the foregoing has been explained
for staple spinning with roller drafting, much of it is equally applicable to toothed
drafting (as in an opening line). Some machines, like cards, have draft ratios of
roughly 100, whereas machines such as drawframes, roving frames, and ring frames
usually have overall drafts of the order of 10. A large number of stages of drafting are
required including those that precede the card.
3.7.7 Effects of roller errors
It is essential that the operating surfaces of all rolls, gears, and other cylindrical
elements should be perfectly round and concentric if periodic errors are to be avoided.
It might be noted that the operating surface of a gear is at its pitch-circle diameter.
An eccentric element produces a sinusoidal error. If a drafting system is left
standing with the pressure acting on the soft cushion rolls, deformations might be
developed in the rubber. Such deformations cause periodic errors in the textile product,
which contains fundamental and harmonic components. Even though an elliptical
roll is a rarity, it is useful to demonstrate the effects. Therefore consider an elliptical
roll in a simple four-roll staple system such as is shown in Fig. 3.14. (Other deformed
rolls will produce somewhat similar effects, irrespective of the type of system.) The
bottom front (delivery) roll has been drawn as excessively elliptical for the purposes
of illustration. All the other rolls are perfectly round and concentric; the back rolls
deliver material at V inches/s. The elliptical bottom front roll rotates at ω radians/s
and the surface velocity is V1 = ω r1 inches/s, where r1 is radius of the roll at the point
of contact. The middle diagram refers to the bottom front roll after it has turned
through 90°. The active radius is now r2 and the velocity is V2 = ω r2 inches/s.
Meanwhile, the back roll speed, V, remains unchanged. Consequently, the draft changes
from V1/V to V2/V as the front roll moves through 90°. As the elliptical roll rotates,
there is a periodic change in draft, which in turn causes a periodic change in linear
density of the output strand. In this case, the periodic wavelength is half the circumference
of the deformed roll. A similar effect would have occurred if the roll had been round
but off-center (i.e. eccentric). In this case, however, the error wavelength would have
been the whole circumference of the deformed roll. Any deformity of the roll produces
an error and, as mentioned earlier, a common cause of such errors is deformation of
the top rolls (which are normally rubber covered). The rubber is used to improve the
grip on the fibers but it is visco-elastic and will deform if the load is left on while the
machine is stationary. It might be added that the rubber coverings harden unevenly
with time and use. The result is that the deformation of the rubber also becomes
uneven. Even if no geometric error is present, an uneven strand is produced because
the rubber deforms in a cyclic fashion. These problems are controlled by using
special tools to measure roundness, concentricity, and rubber hardness on a regular
basis.
There is a further complication. The nip-to-nip distance changes, as shown in Fig.
3.14(c), when an elliptical or any other non-round roll meshes with another. At the
given angle of the bottom front roll, the setting has changed by δL. In effect, there is
a cyclic variation in setting that not only produces a cyclic error of its own but
actually magnifies it. Consequently a great deal of trouble is taken to keep the rolls,
and other elements, round and concentric. The spectrogram is useful in this regard
because out-of-true rolls generate a spike at a wavelength λo, which can be used to
74
Handbook of yarn production
V
V1
r1
(a)
V2
V
r2
(b)
V
V0
δL
L
(c)
Fig. 3.14
Deformed rolls
diagnose the source of the error. Further, any error produced upstream is elongated
by the drafting to be ∆ times as long, where ∆ is the overall draft. Consequently, the
spectrograph can show multiple sources of error. (An actual example is given later,
in Fig. 3.18.)
In symbols:
λo = λ1 × (∆/k)
where λo
λl
∆
k
[3.3]
= error wavelength in strand measured
= circumference of bad roll
= draft between bad roll and point of offtake of the material measured
= a factor which is an integer that takes into account how many lobes
are on the bad roll.
λo and λ1 must have the same units of measurement.
Common principles 75
3.7.8 Drawing a filament
Filaments are made to grip the surface of the drawing elements (godets) by the simple
expedient of wrapping the filaments several times round the godet as shown in Fig.
3.15. The pins, P, lie at an angle; this merely serves to separate the turns on the godet.
The wrap friction effect is the same as is used in a capstan winch; indeed it is
sometimes referred to as capstan friction. Yarn is wrapped round two godets rotating
at different surface velocities, and the draw ratio is calculated from the velocity ratio.
It is important that the surfaces of the godets are concentric with the axis of rotation,
and round, otherwise errors similar to those described earlier will occur. A common
reason for problems arises from irregular deposits of finish and debris on the operating
surfaces.
Vi
Inclined
pins
P
Heat
Filament
flow
P
Vo
Fig. 3.15
Filament drawing
3.7.9 Drawing a sliver (staple processing)
In the drawing or drafting of staple fibers, pairs of rollers are caused to grip the strand
as shown in Fig. 3.16. Weighting by deadweights, springs, or pneumatic systems is
used to press the rollers together and prevent slippage between the fiber and the rolls.
Normally, one roll is made of metal and is fluted; the covering of the other is usually
made of synthetic, elastic material (i.e. it is a cushion roll or ‘cot’). As previously
indicated, the cushion rolls should not be left under pressure, otherwise the rubber
becomes deformed and produces mechanical errors in drafting. Fiber condensers are
necessary to gather the fibers and introduce enough fiber migration to give the sliver
cohesion. Drawframes are made to facilitate easy access to the elements, for example,
76
Handbook of yarn production
Multiple
sliver input
Weighting
Weighting
Rubber-covered
top rolls
Rubber-covered
top rolls
Fluted
bottom
rolls
Reaction
Reaction
Single sliver output
Fig. 3.16 Staple fiber drafting
easy removal of parts liable to fairly rapid wear (such as the cots). They are also
designed to give a direct fiber flow path to minimize chokes.
A sliver is an untwisted rope-like strand of loosely aggregated fibers that are held
together solely by interfiber entanglement. To make good yarn, it is desirable that the
fibers be aligned as well as possible, and this is one of the purposes of drawing.
However, alignment or orientation of the fibers lowers the strength of the sliver.
Sliver becomes weak if it is drawn too much or has too low a linear density. Thus,
there is a limit to how much a sliver can be drawn and there is a limit to how fine it
can be drawn before it is too weak to handle. The minimum linear density is affected
by the degree of fiber orientation and crimp. Therefore, it is normal to set the mechanical
draft to be about the same as the number of slivers fed. This limits the draft for one
‘passage of drawing’. The term ‘passage’ refers to a sliver passing through a drawframe
a single time.
3.8
Consequences of roller errors on the textile product
3.8.1 Periodic errors
Roller or godet defects such as those previously described translate into periodic
errors in yarn, roving, sliver or tow, which are sharply defined. Not only does the
linear density of the material vary in consequence but so also does the structure of the
material strand.
3.8.2 Random errors
Textile strands also contain random errors with a very wide spectrum of errors.
Common principles 77
3.8.3 Cumulative effects of drafting
Where there is a number of drafting stages, the results are cumulative and the range
of error wavelengths can be very large. Yarns show not only an extremely large range
of error but these errors translate into faults in the fabric. The end result of these
irregularities is that the fabrics made from the yarns show undesirable patterning
known as moiré or barré, which reduces their value.
3.9
Control of irregular flow in drawing or drafting
3.9.1 Irregular polymer flow in drawing
An experiment with an undrawn nylon monofilament, or a strip of undrawn or partly
drawn nylon sheet, will show that the draw does not always proceed as expected. The
strand or strip tends to neck as indicated at Fig. 3.17(a) but the process of necking is
not always stable. The thin portion consists of oriented strong material, whereas the
thick portion is largely amorphous and capable of plastic flow. As the draw continues,
material flows from the thick to the thin portion in regions and becomes oriented as
it does so; the flow causes local heating, which tends to localize the flow. A partially
drawn material may have ‘lumps’ in it if several necks form during the draw, as shown
in Fig. 3.17(b); clearly this is undesirable.
Polyester, nylon, acrylic and some others fibers are drawn during normal processing
to improve their molecular orientation, but various materials, such as the cellulosics
have a limited potential for improved molecular orientation by drawing. In the following
discussion, the narrow class of textile materials capable of benefiting from drawing
will be called ‘polymers’ for simplicity even though the term ‘polymer’ really covers
a very much wider range.
The flow of polymer in the drawing operation absorbs energy and the temperature
of the strand rises as it is drawn at high speed. A change in temperature changes the
characteristics of the polymer. To control the mechanical flow, it is necessary to
control the heat flow; hence the use of the hot pin mentioned earlier (see, Fig. 2.13,
p 45). The heat flow and mechanical drag caused by the pin are intended to keep the
neck in its proper position. A polymer has a natural draw ratio, which is a function of
the degree of molecular orientation and the draw becomes unstable if the machine
draw ratio differs from this. The position of the neck will advance or retreat according
to whether the machine draw ratio is lower or higher than the natural draw ratio. If the
machine draw ratio is too high, the tensions rise to the point where the filament
breaks. If it is too low, the system is unstable and the product consists of a mixture
of drawn and undrawn lengths. In a continuous flow process, the position of the neck
has to be stabilized; without such stabilization, the neck is likely to move in one
direction or the other in respect to the godets. (There is an exception when the
Vout
Vin
(a)
Bulge
Vin
Vout
(b)
Fig. 3.17 Unstable polymer flow in ‘neck’
78
Handbook of yarn production
mechanical draw ratio is the same, numerically, as the natural draw ratio.) The neck
retreats or advances until it reaches a godet, where the strand will then break. Any
oscillation of the position of the neck tends to give uneven filaments and therefore
great care has to be taken in the design and operation of the drawing system. This is
especially true if the filaments have to be dyed at a later stage; variations in draw
cause corresponding variations in polymer morphology, which give rise to barré in
fabrics.
Amplitude
3.9.2 Drafting waves in staple systems
The problem of uneven polymer flow in drawing a polymer has its counterpart in
staple spinning. In drafting staple fibers, the effective length of the fiber is an important
factor. Fibers supplied to a drafting system in normal practice vary in length, fineness,
crimp, and finish, natural fibers being more variable than man-made ones. Each of
the variables mentioned alters the force that can be transmitted by a fiber under these
circumstances; the force may be taken as a measure of the effective length of the
fibers. A perfectly smooth fiber behaves as if it were much shorter than it really is.
A crimped fiber is physically shorter than its fully extended length but it engages
neighboring fibers because of the crimps. Thus, the question of fiber length is far
from an easy one. For simplicity, the following discussion will refer solely to the
effects of length variation.
Irregular flow of the fibers changes the position of fibers relative to their expected
positions after drafting and creates unwanted variations in linear density. With natural
fibers, length is quite variable and the error is distributed. The variation produces a
‘hill’ typical of this type of error, as shown in the actual distribution in Fig. 3.18(a)
and it is easily distinguished from a mechanical error (Fig. 3.18(b)), which shows up
as spikes like those at A and B (also there are less easily distinguished peaks such as
shown at C). The first of these types is often called a ‘drafting wave’ and it is a fiberborne error. Fig. 3.18(a) shows two hills, which indicates that two different drafting
waves were created, one from a rearward drafting zone and a large wave from a
forward one. Since a difference between the roll setting and effective fiber length is
an important factor, variations in fiber length can produce undesirable results, which
show up as blotchiness or streakiness in the fabrics.
Actual distribution shown
by bars
Theoretical distribution in gray
Amplitude
0.1
1.0
10
Error wavelength, log scale (yards)
(a) Drafting wave
A
B
C
2
4
8
16
32
Error wavelength, log scale (inches)
(b) Mechanical error
Fig. 3.18
Actual and theoretical variance distributions in a drawframe sliver
Common principles 79
Not only is control of fiber length an important matter, but so is the maintenance
of correct roll settings. There are thousands of drafting zones in a normal mill, and
the aim is to set all of them to standard values appropriate to the material being
processed. Changing the ratch setting (i.e. the distance between the nip lines of
successive pairs of rolls in a drafting system) throughout a mill is a major operation.
Settings of all the drafting systems must be maintained within close tolerances to the
values standard in the particular mill. Also the fiber purchasing agent must seek to
acquire fibers within a standard range of fiber length distributions. These are
management and maintenance problems; changes are not made lightly. Once set up,
the ratch settings are usually maintained until the next maintenance period. Thus, it
is useful to constrain the variability in the fiber population by blending and careful
stock control.
3.9.3 Control of fibers by mechanical restraint
The evenness of the final product, which is usually yarn, can be gravely affected if
these errors just discussed are not kept under control. This is true even when the
faulty drafting is in an earlier process. On the other hand, if the fiber movements
could be constrained to minimize drafting waves, the distribution would approach the
theoretical value as shown in Fig. 3.18(a). In the example, to the left of the picture
there is a significant hill and in the center there is a minor one. The large one came
from the main draft zone and the small one from the break draft (back) zone or an
earlier process. Obviously there must be concern about the irregular flow through the
front draft zone where the draft ratio is the largest.
Fortunately, there are some design features in modern machines that help to restrain
the unwanted relative fiber movement. These features work by adding frictional
forces, which tend to keep floating fibers at a speed at or near that of the back rolls.
These floating fibers within the draft zone are not gripped by either nip. In drawframe
design it is normal to incorporate a pressure bar (Fig. 3.19(a)) or some other device
to restrain these floating fibers. In a ring frame or roving frame, aprons are commonly
used to fulfill a similar function (Fig. 3.19(b)). Aprons are pairs of relatively wide
flexible bands. They are pressed together sufficiently to restrain most fibers so that
they move at the apron speed until the leading ends of the fibers are trapped by the
delivery rolls. The linear speed is usually close to the surface speed of the back roll.
The use of aprons has helped staple spinners achieve remarkable improvements in
yarn quality in this century, by greatly reducing the unstable fiber flow through the
drafting system. The aprons are pressed together by pressure P (Fig. 3.19), merely to
restrain the floating fibers and to allow them to slip without gripping them sufficiently
to cause fiber breakage. This is in contrast to the rollers, which are pressed together
by forces F to eliminate, as far as possible, fiber-to-roll slippage. The aprons press on
the floating fibers and add their influence to the competing effects of the fibers. The
competing forces arise from the frictional contact between the fibers and the front
and back rolls. The concentration of fibers is greatest at the nip of the back rolls and
the surface speed is greatest at the front rolls. As mentioned before, the aprons retard
any premature accelerations of the floating fibers but they must be maintained in
good condition to work properly.
If the draft is too high or the linear density of the strand is too large, the wear rate
of the aprons increases markedly. Aprons are not normally used in sliver drawing
because of the high wear rates. For roving frames it is normal to use long bottom
80
Handbook of yarn production
Pressure
bar
F
F
F
F
V
F
F
(a)
Apron
P
F
F
F
F
F
V
P
F
Apron
(b)
θ
(c)
Fig. 3.19
Fiber control in a drafting system
aprons, such as those shown, to prolong their life, whereas simple short aprons might
suffice in ring spinning. However, many modern ring frames also use long aprons for
the reasons cited. The setting of the forces F and P is critical to good performance.
If F is too small, defects are created in the output strand, whereas if it is too high, the
rubber rolls quickly become damaged. Setting P is part of the art of spinning.
There is a choice of roll layouts but it is usual for the first drafting zone met by the
sliver to be the break draft zone. The second operational zone is the place where the
main draft occurs. Aprons in the break draft zone have been found to wear quickly
and have never gained a foothold in practice. The roll layouts are referred to as ‘3
over 3’ or ‘4 over 4’ according to the number of rolls involved. A 3 over 3 system
implies a system similar to that sketched in Fig. 3.19. A 4 over 4 system has four top
and four bottom rolls but there are still only two drafting zones. The reason cited for
this design by the makers is that it is beneficial to have a rest zone between the two
draft zones, but additionally, the extra spacing insulates one drafting zone from the
worst effects of fiber slippage from the other.
In a conventional ring frame, which has roller drafting, the draft might well be up
to 30, and in certain cases much higher drafts have been used. This applies especially
to air-jet spinning, where the width of the ribbon of fibers presented to the twisting
device is wider than normal. Also, some sliver-to-yarn systems are capable of handling
remarkably high drafts (up to 80). Properly designed high draftframes yield higher
yarn strengths, the yarn is more even, and the productivity is higher than with
conventional frames. Increased precision in setting the high draft machines is required
because high drafts are often associated with increased short-term error.
Common principles 81
Condensers are also used to help control the flowing fibers by forcing them through
an orifice (or other constraint) with a narrow throat, which compresses them.
Compression of the fibers approaching the draft zone keeps the fibers together and
prevents so-called ‘cracking’ of the fiber sheet. It is possible thereby to obtain a
smoother and more even drafting action.
3.9.4 Prevention of fiber slippage over driving surfaces
If the input strands are unequal in size, some may be gripped by the rolls more firmly
than others and the loosely gripped portions may be pulled forward from the feed
prematurely, causing a fault. If the input strands are cored, the outside sheath of fibers
may be improperly gripped by the roller pair. (Cored slivers have a hard central core.
See Section 5.10.4) In such cases, slippage between the rolls and the fiber occurs,
which leads to irregularities in linear density of the output product. A pair of rolls
clamping a strand does not have a simple line contact because the fiber assembly
squashes as shown in Fig. 3.20(a). There is a distribution of pressure in the nip zone,
but a leading fiber end is free from the direct influence of the nip before it arrives at
the compression zone, and the trailing end comes free after it has left. The size of the
compression zone varies according to how thick and compressible the strand is. The
effective ratch setting is thus different from the theoretical value, and the difference
depends on the material being processed. The setting differential just mentioned is a
function of the linear density of the strand being processed and the type of fiber. For
example, the ratch setting in a drawframe should have a greater differential than that
in a roving frame. Another factor also intervenes. The effective length of the fiber
increases during the drafting processes because hooks and fiber-crimps are pulled
out and the fiber is generally straightened. It is therefore not surprising to find, in
practice, that the best ratch settings are often determined by trial and error. If there
is fiber loss or slippage between the fibers and the rolls, the actual draft becomes
slightly reduced.
Perhaps more important is what happens when the amount of slippage varies with
time or position. Obviously, if the slippage varies with time, there are corresponding
variations in linear density of the output strand that constitute a degradation of yarn
quality. Such time-dependent variations can be caused by variations in agglomeration
of the input material. For example, in ring spinning, twist in the roving input tends to
concentrate in the thin spots and this makes so-called ‘tight spots’ more difficult to
draft. The fiber tensions in the draft zone rise and this, in turn, causes slippage
between the fibers and the back rolls. In drawing, the use of slivers of different linear
densities in the creel causes the ‘thinnest’ ones to slip, as demonstrated in Fig.
3.20(b). Strands 1, 3, and 4 are compressed by the forces F and the resistance to
slippage in each of those cases is µF where µ is the coefficient of friction. Strand 2
is too small to be gripped and slips under any applied tension. Even if Strand 2 is
gripped, but to a lesser extent than the others, errors still arise. Highly irregular input
slivers produce similar effects. The same type of condition can often be found in a
comber lap machine in which many parallel slivers are combined side by side and are
drafted to form a comber lap.
Extra control can be obtained by wrapping the fibers partially around any rotating
surfaces with the intention of restraining fibers movement (Fig. 3.19(c)). A partial
wrap design improves the grip between the back rolls and the flowing material. The
wrap idea is also sometimes applied at the front rolls and permits some control of the
82
Handbook of yarn production
Strand
Roll
Compression zone
(a)
F
F
F
Roll
1
F
2
Roll
3
4
F
F
(b)
Fig. 3.20
Unequal compression of strands in a nip
fibers passing through the twist triangle on the output side. Wrapping a strand around
a roll to get a better grip is common to both staple and filament processing.
In filament processing, the use of godets to grip the filaments is very common. A
pair of rolls squeezed together on an incompletely solidified filament would cause
flats on the surface, which would be undesirable under most circumstances. The use
of a number of wraps around a large diameter godet roll produces much less surface
stress for a given gripping power. There is a similarity in concept between this and the
partial wraps just mentioned. However, when multiple wraps are used, it is necessary
to keep the coils separated. This is achieved in a very elegant fashion by the inclined
pins shown in Fig. 3.15. Also in the godet system, the resistance to slippage is a
function of the cumulative wrap and the coefficient of friction (µ) between the fiber
and the metal. In these cases it is important to ensure that an adequate number of
wraps is used and that the surface of the godet roll is free from contaminants, which
would affect µ. In the production of filament yarn, each spinning head produces only
one yarn. However, when tow is being produced, there is more ambiguity in the load
distribution between the component feeds and consequently there is a greater chance
of error. The filament length is virtually infinite and problems associated with
longitudinal fiber migration do not arise.
Common principles 83
3.10
Doubling
3.10.1 The principle of doubling
If m similar streams of fiber converge, the theoretical variance of the total combined
stream is theoretically reduced to 1/m of that of the individual input streams. The
theoretical coefficient of variation (CV) is reduced by a factor of √(1/m), where CV
= St dev/mean and St dev = √Variance. The same applies if m similar strands of sliver
(or other strands) are laid in parallel in the feed to a machine. This combining of
multiple fiber streams is known as doubling. An example is where slivers are placed
in the creel of a drawframe; they are combined into one during and after the drawing
phase of the operation. (A ‘creel’ is that portion of a machine where a multiplicity of
input strands are removed from their packages and are delivered to the strand processing
device. An example is shown later in Fig. 6.1.)
Other cases exist of placing streams of fiber or strands in parallel and combining
them. Many errors are created in processing; consequently, even if the material is
delivered to the process evenly, the output contains variation. Sometimes doubling is
used to offset the errors and the theory works reasonably well providing the errors are
random. (Strictly speaking, the mean values and variances of each strand should be
similar for the theory to apply.) The actual CVs are higher than the theoretical values.
When doubling and drawing are combined, the input materials are doubled to
reduce the long-term errors; however, new errors of shorter wavelengths are added as
a result of the process of elongation. There is an exchange of relatively long-term for
short-term error.
3.10.2 Combinations of drawing or drafting with condensation
Doubling occurs in some processes that are not widely regarded as a form of doubling.
For example, in filament tow production, parallel streams of filaments are combined
before they are chopped or otherwise separated into staple fiber. The multitude of
parallel streams reduces the total error in the output material. The collection of the
cut material also creates a further doubling effect because of the mixing that takes
place as the stream of cut fibers is condensed into the bales. In condensation, deceleration
causes the incoming product stream to fold and produces multiple doubling of adjacent
elements. The effect is a reduction in variance which, although it is somewhat less
than the theoretical value, is still substantial.
Drafting in the early stages of staple fiber processing is always followed by a
condensation stage because of the need to control the flow of several machines in
series. Doubling occurs in every condenser, chute feed, blending machine or any
other place where the fiber flow rate is slowed down and the fibers accumulate. In
fact, if it were not for this doubling, the drafting problems in the opening and carding
would be noticed more. As it is, the increases in variance caused by the toothed
drafting in the opening line are mostly offset by the large doubling factors that also
prevail. This is not to say that the extra variance did not exist.
3.10.3 Homogenizing multiple streams of a nominally
similar product
Another valuable aspect of doubling is that it can be used to offset variabilities from
one set of equipment to another. Each set of equipment tends to produce a product
84
Handbook of yarn production
with different characteristics according to its state of maintenance, setting, and design.
If a downstream machine were always fed from the same source, there could be
differences in product that would show up in the fabric, in the form of barré. This is
called channeling. The effects are greatly reduced if the product is properly distributed
to all the downstream machines in a systematic way that avoids patterning.
According to Uster [2,3], the CVs of linear densities of sliver from the card, 1st
drawframe, and 2nd drawframe sliver for the top 10% of yarn makers were 3.4%,
3.5%, and 3.6%, respectively in 1982. For the bottom 10%, the respective CVs were
5.7%, 7.5%, and 7.2%. This means that the gains due to doubling at the drawframe
are just about offset by the losses in regularity caused by drawing. One purpose of
doubling is to blend product streams. The implication is that it is not just the linear
density of the product stream that is important, but so are all the other aspects of
homogeneity of the stream. In the case of filament processing, this might be represented
by differences in polymer morphology. In staple processing, it might be represented
by differences in fiber attributes other than linear density. Some practical results for
cotton processing are shown in Table 3.2. In this case, the CVs of short fiber content
after carding rose slightly when compared with the average value in the appropriate
bale slice, whereas the CV of fiber fineness dropped considerably.
3.11
Effects of shear
3.11.1 Definition of shear
Shear may be defined as trapezoidal deformation or relative movement of elements
in a structure, which causes the elements to slide over one another. When a viscoelastic body is stressed, some dislocations in the structure remain after the external
stress is removed. There is often a residual stress pattern too. Such dislocations
and residual stresses form the basis of a number of phenomena that are exploited
in yarn manufacturing. They can be considered at the molecular and at the fiber
levels, the former relating mostly to texturing and the latter mostly to staple yarn
processing.
At the molecular level, the elements are segments of long-chain molecules; the
movements are measured in microscopic units and it seems suitable to describe these
movements as dislocations. In many solid materials, crystalline areas are separated
by amorphous areas and the ‘crystals’ can, and do, move with respect to one another
under stress.
In staple fiber processing, fibers migrate and the composition at any cross-section
of the material undergoing drawing changes because of that processing. The phenomenon
can still be regarded as visco-elastic but obviously the elastic forces are proportionately
less than those in polymer molecular dislocations. The relative movements of fibers
Table 3.2
Bale slice
Sliver
Fiber attributes
Linear
density
%CV
Fiber fineness
micronaire
%CV
Short fiber
content
%CV
–
3.1
4.2
2.2
15.5
17.3
Common principles 85
can be quite large and, historically, the term migration has been applied; therefore it
is proposed to extend that practice here in respect of other fiber movements.
3.11.2 Molecular dislocations
The visco-elastic properties of polymers change when heated, especially at the transition
temperatures, i.e. softening point (Tg) and melting point (Tm). Above Tg, many of the
bonds between molecules are broken and relative movement of segments of the
molecules occurs. When the temperature drops below Tg, new bonds are made with
the molecules in their new shapes and relative positions. This makes possible the heat
setting of filaments into desired yarn textures suitable for commercial use.
When two polymers are involved, as in bicomponent yarns, differential stress
patterns caused by the cooling polymer cause filaments to curl, coil, or loop. Also,
the visco-elastic constants of a layer along the length of yarn can be altered to
produce similar effects. The edge-crimp method is one such example, where the
filament is dragged over an edge to produce a disoriented layer and that is sufficient
to make the fiber deform significantly (see Chapter 4).
Similar behavior can be found in staple yarn processing, where the result is not at
all desirable. Fibers are dragged over sharp teeth in a number of machines and they
stand a chance of becoming edge-crimped. Fine fibers that are stressed by this kind
of action can form into tiny tight balls called neps (which are a cause of loss of
quality).
3.11.3 Lateral fiber migration
At the fiber level, most of the phenomena discussed relate to relative movement of
one fiber with respect to its neighbors. For example, in ring spinning, segments of
some fibers are subject to higher stresses than others in the twist triangle, and they
move radially within the structure of the yarn. Highly tensioned fibers tend to move
to the core and slack fibers move to the outer perimeter. In consequence, fibers thread
their way between what would otherwise be concentric layers and stabilize the structure
to make it self-locking. This is called fiber migration but it really should be called
lateral fiber migration or some such term. In air-jet texturing, segments of filaments
are forced across the yarn structure so that they, too, form an interlocked stable
structure. Interfiber friction is the medium by which the structures become locked. A
useful analogy to these effects is the common knot.
3.11.4 Longitudinal fiber migration
As explained in Appendix 8, fibers flowing through a draft zone do so irregularly;
there is shear between the fibers and some have a higher mean velocity through the
drafting system than others. Generally, shorter fibers accelerate from the nip of the
back rolls in a drafting zone earlier than longer ones that enter at the same time. The
relative motion between two fibers delivered by the device is called ‘longitudinal
fiber migration’.
Maximum longitudinal fiber migration = (∆ – 1) × (L – S)
[3.4]
where ∆ is the draft, L is the length of the long fiber, and S is the length of the short
fiber.
86
Handbook of yarn production
The ratch setting is usually set a little larger than L and the longitudinal migration
of fibers varies between zero and the amount given in Equation (3.4). The fiber
population of a given cross-section of the entering material differs from the population
of the corresponding zone in the emerging material; it becomes difficult to predict
the performance of the fibers in drafting because of these changes in fiber population.
3.11.5 Effect of migration on evenness
The longitudinal fiber migration just described has an important effect on the evenness
of the strand. To demonstrate the mechanism by which the evenness is degraded,
consider the following example. The strand shown in Fig. 3.21(a) has two components
and the composite strand has been leveled by some mechanism such that the evenness
of the composite is perfect and the linear density of it is constant at nav units. However,
component B has a thick spot, which is compensated by a conforming thin spot in
component A. A normal autoleveling device cannot discriminate between the relative
natures of the two components and can only make its adjustments based on the total
linear density. Two conjugate intervals are marked by hollow headed arrows. The
component A is now moved to the right by some means. Figure 3.21(b) shows how
the linear density changes near the blend anomaly. The marked intervals do not
change their heights but they become separated. Bearing in mind that the height of
the diagram represents linear density, it will be observed that the strand is now
uneven. The linear density now ranges within the limits nav ± n1. Variations in fiber
length along the strand are converted into variations in linear density. If the concentration
of short fibers varies (as was shown earlier), then the longitudinal fiber migration
varies and the result is that a perfectly leveled strand can become uneven after
passing through the process stage. Thus, for example, a roller drafting system converts
variations in fiber content into variation in linear density. It can be seen that irregular
blending produces some undesirable side effects.
3.12
Integration of sub-processes
3.12.1 Historical examples of process integration
As will be detailed more exactly in Chapter 5, the development of cotton card sliver
production in the twentieth century provides an example of the integration of a
variety of machines arranged in serial fashion in a process line. This is the first case.
A
nav
B
(a)
n1
A
B
–n1
(b)
Fig. 3.21
Longitudinal fiber migration and strand evenness
Common principles 87
In the 1950s, each machine was free standing. The fiber transfers, to and from the
machine, were executed by manual labor and were also controlled manually. By the
close of the century, the transfers were automatic and so was the control. Very little
labor is involved in the operation of the series of machines that open, clean, and blend
the fibers taken from the bale. No appreciable labor is needed until the sliver emerges
from the card. In contrast, the subsequent stages (in which the card sliver is converted
into yarn) still involve considerable manual work. This is the second case. Various
schemes of automation are being applied, but not universally. The first case shows a
mature system and the second one shows a system developing in a somewhat similar
direction. Integration and automation of process lines are a theme common to all
phases of yarn production.
3.12.2 The driving force behind process integration
Competitive pressure mandates economical means of production. Since labor costs
have been much of the total cost of yarn, there has been a powerful motive to reduce
the amount of labor needed. This pressure is offset by the costs of capital needed to
implement technical solutions. Consequently, not every solution is adopted by industry
and those that are adopted are put in place with considerable caution. In the first case
just cited, the solutions adopted involved relatively modest amounts of capital and
yet, over the years, have yielded significant reductions in the labor costs involved.
The linking and automation of the next series of processes require investment and the
savings are relatively modest; thus progress is relatively slow. Another aspect of
reducing labor costs in spinning is the need to deal with end-breakages. There is a
contrast between the initial and final processes that helps explain the dilemma. In the
blow room there are normally only two or three production streams, whereas in
spinning there are many, many parallel streams. Consequently, the acceptable capital
cost per stream is relatively low in the last case and this hinders the development of
process integration. Nevertheless, there is movement in this direction and future
generations will see it mature.
Filament yarn production seems to be less affected because of the shortness of the
process line, but it is not immune to the pressures. For example, extrusion, texturing,
and drawing were originally separate operations in series, but now draw-texturing is
well established with little or no increase in capital cost. The relatively small labor
cost was reduced. The development of the use of partially oriented yarn (POY) was
responsible for that commercial advance, rather than some mechanical solution. Thus
it can be seen that there is no single route to increased process integration, but the
economic pressures ensure that progress in that direction will continue.
References
1.
2.
3.
Lorenz, R R C. Yarn Twisting, Text Prog, 16, 1/2, 1987.
Anon. Uster Bulletin, Zellweger Uster Inc, Uster, Switzerland, No 31 Dec 1982.
Furter. R. Uster Bulletin No 36, Zellweger Uster Inc, Uster, Switzerland, Oct 1989.
4
Filament yarn production
4.1
Introduction
Although most filament yarns used today are synthetic fibers that need texturing,
there are some that need no modification in this way. Industrial filaments made from
synthetic polymers constitute one case and natural filaments, such as silk, another.
This chapter will concentrate mostly on textured yarns but a brief discussion of
silk throwing will be included for the sake of completeness, at the end of the chapter.
Industrial filaments are so diverse that little discussion will be given. Suffice it to say
that the majority of the successful processes exploit the exceptional strength that can
be obtained with some drawn polymers.
During the period since 1975, manufacturing facilities have sprung up in countries
such as China, Taiwan, Korea, Mexico, and Brazil. These countries operate to fill
some of the demand of new markets. They also serve the established ones in the
USA, Japan, Europe, and other developed areas. Such changes affect the price and
distribution of the materials. The total consumption of textured yarn in the USA,
Japan, and Europe has declined but there has been steady growth in industrial and
carpet yarns. According to Wilson and Kollu [1], 51% of the textured yarn produced
in 1983–4 was false twisted polyester filament, 22% was false twisted nylon, 18%
was bulked continuous filament (nylon and polypropylene), and the remainder was
made up of air-jet and other forms of textured yarns. Obviously, false twisting is very
important in this field. However, the market has forced many filament yarn makers
to move to products nearer to staple yarns in character and consequently the use of
air-jet texturing has risen. Atkinson and Wheeler [2] state that air-jet textured yarns
have maintained about 5% of the market for false twist textured yarns and most of
that goes into automotive upholstery. Polyester has largely displaced nylon in that
particular market.
Filament yarn production 89
4.2
Texturing filament yarns
4.2.1 Purposes of texturing
The prime purpose of texturing filament yarn is to create a bulky structure that is
desirable for the following reasons:
1
2
3
4
5
The voids in the structure cause the material to have good insulation properties.
The voids in the structure change the density of the material (which makes
possible a lightweight yarn with good covering properties).
The disorganized (or less organized) surface of the yarn gives dispersed light
reflections, which, in turn, give a desirable matte appearance.
The sponge-like structure feels softer than a lean twisted ‘flat’ yarn.
The crimped or coiled filament structure gives a lower effective modulus of
elasticity to the structure when compared with that of a flat yarn.
From this it will be realized that, in order to make yarns to these specifications, it is
necessary to deform the individual filaments and set, or otherwise hold, them in the
desired deformed condition. When deformed in this way, the filaments in the whole
bundle are unable to lie side by side in close contact and the required voids are
produced.
Furthermore, the non-straight, separated filaments are much more easily deformed
than are those in a flat yarn, and one obtains a softer hand and greater ‘stretch’. There
are two general classes of textured yarns that relate respectively to thermoplastic
yarns only and to those which can be more widely used.
In general, the first classification involves the stages of deforming, heating, cooling,
and relaxing the filaments. The process is known as heat setting despite the fact that
it is the cooling that does the setting. Theoretical filament structures are shown
diagrammatically in Fig. 4.1.
In the second case, the texturing of non-thermoplastic materials, filaments are
deformed and are held in their deformed state by frictional contact with the neighboring
filaments. An example of the latter is the air-jet method that will be described later
in this chapter. Meanwhile, we will continue with heat set yarns.
4.2.2 Physical basis of texturing
Before considering the methods of false twisting, let us review the mechanics involved.
It will be recalled that the process phases in false twist texturing consist of:
(a)
Fig. 4.1
(b)
Theoretical yarn structures
90
Handbook of yarn production
1
2
Deforming the filaments.
Applying heat to raise the filament temperature above the glass transition
temperature, Tg.
Cooling the filaments to below Tg.
Rearranging the filaments under suitable tension.
Winding the textured yarn.
3
4
5
Theoretically, phases (1) and (2) can be interchanged or be coincident, provided the
deformation persists until the filaments are cooled below Tg and the polymer becomes
set. However, time is a factor in determining the degree of set achieved and, in high
speed machinery, it is usual to apply heat as soon as possible in the process. If
temperatures of some polymers are raised too high, they tend to yellow and this gives
trouble with the end products, particularly those of light color shades. The deformation
can be of any kind, but in false or real twisting, the primary modes of deformation are
torsion and bending. Since the real twist process is simple, it will be used for explanation
although it is no longer commercially important.
4.3
Real twist texturing
Explanations are a little easier if we consider the early types of discontinuous processes.
Various forms of twister were used to induce the initial deformation. A batch of
packages of yarn was then taken from the twister and placed in an autoclave.1 The
temperature of the yarn was raised above Tg (but below Tm), and then allowed to cool.
The product taken from the autoclave was non-twist lively or ‘dead’ (see Fig. 3.4), but
the fiber deformations were set into their newly twisted shapes. To develop the bulk,
it was necessary to untwist the yarns until the filaments were approximately parallel
and separated, and then relax them. It will be noted that filament separation in the
phase (4) was necessary for the bulk to form without undue interference between
neighboring filaments.
In untwisting yarn from the set condition, a torque is applied to each filament. The
sum of the individual torques is the total applied to the yarn. The torque places it in
a state of stress, which is retained until the fibers are relaxed. Untwisting and relaxing
the yarn allow the newly imposed stresses to be relieved by changes in the shape of
the filaments as they move within the structure during the process of relaxation. This
form of texturing is shown diagrammatically in Fig. 4.2. When relaxed, each filament
seeks a minimum energy state, two of which are depicted in Fig. 4.1. If the structure
is open enough, most of the filaments will achieve one of the minimum energy
shapes, but a tight structure prevents full relaxation. In the latter case, not all the
potential bulk is developed. A normal yarn structure will consist of shapes similar to
those shown, or combinations of them if yarn is untwisted and the filaments are
separated before release. Some methods of texturing produce alternating directions
of coiling. The result is that the yarn produced has little or no twist liveliness because
torques from the opposing filament coils cancel. This form of texturing is shown
diagrammatically in Fig. 4.2.
Consider extreme cases. The adjacent helical coils in Fig. 4.1(a) take up a great
1 A vessel that uses high pressure steam to obtain the necessary temperatures. For the characteristics
of steam, see Appendix 3.
Filament yarn production 91
Twist
Heat
Cool
Untextured
filaments
Untwist
Separate and relax
Textured
filaments
Fig. 4.2
Principle of twist texturing
deal of space and we have a so-called ‘bulky’ yarn. The other model, Fig. 4.1(b),
consumes relatively little space and we have a low bulk, high stretch yarn. As the yarn
is extended, the intermittently snarled filaments are progressively converted to straight
parallel filaments. There is a great deal of yarn stored in the snarls, and, consequently,
there is a surprisingly large extension of the yarn before the snarls are fully converted
to straight parallel filaments. Furthermore, the tension needed to pull out the snarls
is relatively low, and thus the yarn behaves as a low modulus material (until all the
snarls are removed). Of course, as the filaments change from the snarled to the
straight condition, they are subjected to torsional and bending stresses, and energy is
stored in the extended yarn. Once the tension is removed, the yarn attempts to return
to a minimum energy state and contracts. Thus, the stretch yarn behaves rather like
a rubber band and its principal characteristic is the enormous and almost elastic
extension that becomes possible. A practical yarn is intermediate between the extremes.
There are varying proportions of each kind of minimum energy shape according to
the method and conditions of texturing. Also, there are modifying factors. Helical
portions tend to intermesh, parallel portions tend to migrate (and become non-parallel),
many filaments fail to reach their minimum energy state, and many filaments interfere
with one another. Consequently, there is a wide range of combinations of bulk and
stretch that can be achieved, but generally the higher the stretch capability, the lower
the bulk. Of course, even the adjacent coil model provides a yarn with a moderate
degree of stretch because the helices act as coil springs. In practice, the breaking
elongation might vary from 10% for a bulked yarn to 500% for a stretch yarn.
92
Handbook of yarn production
4.4
False twist texturing
4.4.1 General comment
One of the most important types of yarn modification is false twist texturing. As
mentioned in the last chapter, a running yarn twisted as shown in Fig. 4.3 causes false
twist to be trapped between the feed system and the twister. The feed yarn has little
or no twist, the yarn between A and B has false twist, and the yarn leaving B has the
same twist as the input. If heat is applied in the zone AX and the yarn is cooled in
zone XB, then the yarn approaching B will be heat set in the twisted condition.
Overfeeding (not shown) and untwisting slackened filaments at B facilitates the
necessary fiber rearrangement and separation. (An overfeed is where the input speed
is slightly more than the output speed.) When the filaments relax, the uneven contraction
of the filaments causes them to rearrange themselves laterally. If heat is applied in
zone CD, the latent crimp can be developed to produce a bulked, set yarn in one
continuous process.2 In the particular case shown, a godet is used to grip and feed the
input yarn; however, no twister is shown for reasons of clarity. All the phases mentioned
in the previous section are embodied in this continuous process. The integration
reduces costs of machinery and material transportation. The savings have been so
large that false twist texturing has become a major system for yarn production. The
Untextured yarn input
Godet
A
False
twisted
yarn
Heat
X
B
Cool
Twist
Zero twist
filament
output
Develop
texture
Fig. 4.3
C
The temperature
in the zone AX is
raised above Tg
D
False twist texturing
2 Notice that care was taken to avoid saying that the output yarn had no twist.
Filament yarn production 93
means of twisting has changed and the systems will now be reviewed in a more or
less historical sequence.
4.4.2 Pin twister type of false twist texturing machines
To heat set the twisted filaments and relax them afterwards to produce bulk, it is
necessary to heat the running filaments at two places and so we have two-heater
machines to produce the developed yarns. To produce yarns in which the filaments
have not been relaxed only one heater is required. Examples of a two-heater machine
are shown to a small scale in Fig. 4.4. It is necessary to use high twist levels to
produce adequately textured yarns; for example, with a 70 denier yarn, one might
well use some 80 tpi. (This would give a TM of about 10 on the cotton system.) To
get high production, it is necessary to use very high twisting speeds, of around
500 000 r/min. This calls for special designs of twisting unit in which the mass and
size of the rotating element are as small as practical (or the element is eliminated).
It also calls for special bearings, or suspension systems. In the pin twister shown in
Tensioners
Godets
Heaters in
false twist
zone
Twisters
Twister pin
(enlarged)
Secondary
heaters
Godets
Winders
Take-up
packages
(a)
Fig. 4.4
(b)
Manufacture of false twist yarns
94
Handbook of yarn production
Fig. 4.4, the spindle is frequently less than 0.25 inch diameter × 1.5 inch long
(approximately 6.4 mm diameter × 38 mm) and it is held against drive rollers by a
magnetic field; this obviates the need for a direct bearing. The bearings of the drive
rollers have to rotate at only a fraction of the speed of the spindle (typically 12–15%).
It should be noted, however, that the spindle gets very hot because of air drag and
magnetically induced eddy currents within the metal. Also, the false twist pin (shown
inset) is usually made of ceramic or sapphire to withstand the abrasion caused by the
yarn passing over it.
A given element of polymer must reside in the hot environment for a sufficient
period to reach Tg because it takes time to soften the polymer. If, for example, the
time is 0.5 second, the spindle speed is 500 000 r/min and the twist is 80 tpi, the
heater length has to be at least 52 inches. Thus it can be seen that the heaters must be
long.
It also takes a significant time for the yarn to cool sufficiently to freeze it into the
twisted configuration. Thus, a certain distance is needed between the heater and the
false twist pin. The needed heating and cooling lengths increase with spindle speed
and this leads to increases in the threadline length. Not only do high production
machines become very tall, but there is also increasing difficulty in handling the
long, heated filaments. Frictional drag of the yarn over the heater plate is a significant
factor. The frictional coefficient is modified by the fact that the yarn rotates at high
speed about its axis as it passes over the heater plate. At very high speeds, the design
of the heater becomes extremely important and it sometimes becomes necessary to
use forced cooling of the yarn leaving the heater.
Where two heaters are used (to produce a set yarn), the threadline length is almost
doubled, as shown in Fig. 4.5. If the threadline is vertical and the two heaters are
immediately above one another, a two-story building becomes necessary for high
Feed roll
Oiler
Feed roll
Second
heater
Winder
Feed roll
False
twister
Tensioner
First
heater
Floor
Feed roll
Fig. 4.5 Two-heater false twist machine
Filament yarn production 95
speed machines. Alternatively, a more complex threadline may be used; for example,
the heaters might be inclined to the vertical. In all cases, the modern machines need
a great deal of headroom. Threading up (or ‘stringing up’) needs skill because of
difficulties in handling the hot, high speed yarns. It might be added that the use of air
to piece and to thread godets, and other high speed elements, is very common in the
filament industry.
To reiterate, the temperature of the polymer has to be raised to a level between Tg
and Tm. Within these limits, the higher the temperature, the better the set, but as the
temperature approaches Tm, the yarn strength deteriorates and excessive differences
in dye affinity are likely to be created. Atmospheric conditions should be controlled
because moisture affects the setting process and can lead to degradation of the polymer.
Generally, an air temperature of 75 ± 5°F (24 ± 3°C) and an rh of 65 ± 2% are used,
but the conditions might vary according to the yarn being textured. Excessive humidity
causes yarn to drag over contact surfaces, which leads to erratic tensions in the yarn.
This, in turn, leads to variations in the bulk developed. Insufficient humidity leads to
the production of static electricity and, on all of these accounts, control is very
important.
Tension in the yarn within the heater is controlled by the feed uptake rates. The
feed rolls have to be adjusted to give an overfeed of 2 or 3% to take into account twist
contraction and shrinkage. Insufficient overfeed leads to high tension, which causes
unacceptably high end-breakage levels and low bulk. Too much overfeed leads to low
tension, which results in the formation of tight spots (sometimes called ‘voids’), poor
set, and, again, deterioration in the end-breakage or filamentation rates. The tight
spots are seen as apparently untextured (or lightly textured) segments in the yarn that
show up as defects in the fabric. These tight spots are caused by twist slipping over
the false twist pin in an erratic manner. Segments of yarn leave the twist pin containing
real twist; a twisted segment of yarn is unable fully to develop bulk. Over-twisting
the yarn can produce a similar result. The twist level determines the hand and appearance
of the material; a high twist gives the fabric a soft, fine texture, whereas a low twist
yields a rough, pebbly look. High twist gives a relatively high crimp contraction and
therefore more stretch potential. It also causes more tight spots and weakens the yarn
(up to 20–30% strength loss for nylon, but very little for polyester or acetate).
Fiber producers apply a finish to the surface of the filaments immediately after
extrusion to help drawing and subsequent operations. The finish is intended to reduce
static electrification and friction, but when it is heated in the texturing operation, any
volatile fractions of the finish are driven off, giving rise to unwanted fumes. Heavier
fractions can oxidize or otherwise deteriorate and cause problems with the deposit of
solids in the heater zones. This is especially so if high heater temperatures are used
(say 400°F, about 200°C). Loss of the fiber finish can also create a problem and it is
often desirable to apply a lubricant after texturing. These so-called ‘coning oils’
replace the losses and facilitate winding and fabric manufacture. However, any such
oil should be stable and capable of being scoured away without detriment to the color
or performance of the yarn. A sufficiency of fiber finish or additive is important but
excessive amounts of finish are to be avoided. Also, variations in the add-on levels of
finish should be kept to a minimum.
Some fibers are dulled by the addition of titanium dioxide (TiO2); this additive
affects the wear rate of guides and pins. Such wear can adversely affect the quality of
yarn being produced as well as the efficiency of the operation.
With a single-heater machine, it is necessary to soft-wind the yarn packages to
96
Handbook of yarn production
permit satisfactory subsequent autoclaving to produce set yarns. With two-heater
machines, it is necessary to overfeed the yarn into the second heater to allow the
crimp to develop. This overfeed level is normally about 4 to 5%. The single-heater
machine used in conjunction with an autoclave is less efficient than a two-heater
machine. With the batch process of autoclave setting, variations between batches are
more likely and thus there is an increased risk of producing barré in the fabrics. This
is because of the changes in bulk and dye affinity arising from non-constant heat
treatment conditions. Whatever system is used, great effort has to be taken to strictly
control all temperatures, tensions, and twist levels so that they are similar from
spindle to spindle, from time to time, and from batch to batch. The consequence of
a failure to control, in all these respects, is that streaks and barré will be produced in
the dyed fabric. Modern machines are equipped with control devices; in addition,
strict quality control is exercised by means of proper sampling and testing. However,
the potential flaws are rarely visible in the yarn coming from the machines. Therefore,
it is necessary to carry out tests on dyed yarn at a very early stage before large
inventories are accumulated.
4.4.3 Limitations of the pin twister machine
The size of the false twist spindle dictates the maximum rotational speed that can be
used. Remembering that the power absorbed by a spindle due to air drag alone is
roughly proportional to D4U3 (where D is the diameter and U is the rotational speed),
it will be readily realized that the spindle has to be kept as small as possible (see Fig.
4.6). However, there is a practical limit to smallness. It must be possible for a knot to
pass through the spindle and this means that the diameter of the central hole in the
spindle must be several times that of the yarn diameter. Thus, with 150 denier (167
dtex) yarn, the central hole must be of at least 1 mm (≈ 0.04 inch) diameter; for
heavier yarns, the hole must be larger. Requirements for the false twist pin and the
need for sufficient space to permit the threading operation control the minimum size
of the largest diameter of the spindle.
Centrifugal forces acting on the yarn, spindle and drive system can be very high.
In the case of the spindle, it is necessary to ensure that it is dynamically balanced;
otherwise, at high speeds, it will tend to ‘tramp’ like an unbalanced wheel on a car,
and the drive tires might suffer considerable damage as a consequence. As well as
encountering considerable centrifugal force, these tires are also subjected to high
temperatures (due to frictional heating). The combination of the two can cause polymer
creep, with a result that the tires sometimes grow in diameter during service. A
change in diameter alters the forces acting on the surfaces. Growth usually signals
impending failure of the tires. The surface of the tires can also suffer damage due to
high shear stresses caused by the localized loading, and the damage shows up as a
pitting of the surface. If the spindle is unbalanced, the loads are greatly increased and
failure of the tire surface is hastened. There is usually a finite life for these tires and
the units have to be replaced from time to time. Damage and imbalance cause an
increase in noise level and faulty machines are difficult (if not impossible) to operate
within the legal noise level limits of some countries.
The yarn is pressed against the wall of the axial hole inside the spindle by the
centrifugal forces. This causes the yarn to drag, which can cause filament breaks, and
since the drag is related to ω2d (where ω is the spindle speed and d is the hole
diameter), it is obvious that a large central hole in a very high speed spindle is
Filament yarn production 97
Section X–X
Yarn lifts
off pin
Y
Y
Yarn presses
against wall
Hard pin
(a)
Access hole
Section Y–Y
X
X
(b)
Fig. 4.6
A pin twisting element
undesirable. This is especially important when producing fine yarns. Eccentricity can
induce quite strong yarn ballooning in the heater zone. As will be realized, the
variations in distance between the yarn and the heater surface can greatly affect the
local heat transfer rate. Under certain circumstances, this can affect the set of yarn in
a periodic fashion and produce patterning or barré in the final fabric.
Additionally, centrifugal force acts on the yarn wrapped around the pin inside the
spindle. A portion of the yarn wrap sometimes moves away from the pin as shown in
the enlarged sketch in Fig. 4.6(a). Eccentricity of the wrap causes it to pull away even
more and the eventual restraint is from the walls of the access hole. The grip on the
yarn by the pin is then reduced and twist slips over the pin. Intermittent slippage of
this sort generates undesirable tight spots in the yarn. Twist is associated with tension
and this is an unstable relationship, which can lead to surges that give operational
problems as well as the undesirable periodic tight spots.
At the high linear speeds of yarn take-up associated with high speed operation,
there is frictional heating of some of the outer filaments of the yarn. Such heating
occurs (a) at the twist pin, (b) in the central hole of the spindle, (c) at various guides,
and perhaps (d) at the heater surface (if the yarn is not properly controlled). At these
‘hot’ spots, there is likely to be filament damage or breakage. The undesirability of
98
Handbook of yarn production
breakage has already been mentioned. Apart from the problems of wild filaments
(uncontrolled filaments not bound into the body of the yarn) and reduced yarn strength,
the local overheating might cause segments of some filaments to fuse together.
Furthermore, it might result in changed local yarn extension, or it might change
dyeability at the local spots. Whichever combination of such faults is generated, it
impairs both the efficiency of the operation and the quality of the product. In all these
cases, the higher the speed, the worse the problems become. Consequently, there must
be practical upper limits to speed and this, in turn, means that there are practical
upper limits to the productivity of pin twisters. Improvements in the technology
continue to raise the limits, but it becomes increasingly more difficult and costly to
do so. In fact, the rise of friction twisting caused further machinery developments of
pin twisters to show unsatisfactory returns on investment. Whether pin twisters will
find a market in the future is uncertain.
4.4.4 Friction twisters
In the search for ever higher productivity, the false twist element has, over the years,
become ever smaller. The ultimate stage was that the diameter of the high speed
rotating element was reduced to that of the yarn itself. After that we had friction
twisting with its enormous potential for increased speeds. An example of friction
twisting is shown in Fig. 4.4(b) and two embodiments of the principle are shown in
Fig. 4.7.
In Fig. 4.7(a), friction between the bore of the rotating tube (bush) and the yarn
causes twist to be inserted into the yarn. In Fig. 4.7(b), it is the friction between the
outside surface of the disk and the yarn that gives the effect. In both cases, there is
slippage and therefore it is not possible to calculate the twist insertion rate from the
ratio of diameters (i.e. rotating element diameter/yarn diameter). It is better to consider
the torque generated. From Fig. 4.7(a), it may be seen that the reaction F must
balance components of yarn tensions Tin and Tout resolved in a direction perpendicular
to the axis of the bush. For the present purpose we may ignore the components F3 and
F4. In other words:
F = F1 + F 2
[4.1]
where F1 = Tin cos γ
F2 = Tout cos α
Since torque is (force) × (radius of action), and the relevant radius is that of the
yarn under operating conditions, we may write:
Torque = µkFd/2
[4.2]
where d is the diameter of the yarn in the free state, and k is a factor that takes into
account the local compression at the contact zone between it and the twister, as well
as the end effects at the edges of the twister. The factor k < 1 and µ is the coefficient
of friction. In the simple case shown in Fig. 4.7(a):
Torque generated by the twister = µ (kd/2)(Tincos γ + Toutcos α)
[4.3]
If n is the linear density of the yarn, the effective yarn radius is K√n, where the factor
K includes k / 2 used in equation (4.3) as well as the factor relating diameter to linear
density:
Filament yarn production 99
Tout
F3
F2
Vout
α
U r/m
γ
F1
Vin
F
F4
Tin
(a)
Rubber end caps
β
F
Vout
α
Tout
γ
Tin
Vin
(b)
Rubber tire
(c)
A selection of disk profiles
Fig. 4.7 Friction twister elements
Torque generated by the twister = µK√n(Tincos γ + Toutcos α)
[4.4]
In other words, the torque is influenced by the linear density of the yarn and its
compressibility. It is also influenced by the coefficient of friction, the tensions applied
as well as the angles taken up by the entering and departing yarns.
Similar logic can be applied to the disk twister, but in this case, K is further
affected by the attitude of the yarn on the surface of the disk (the angle β shown in
Fig. 4.7(b)), which is discussed in the following paragraphs. The disk type of machine
is more widely used, therefore we shall restrict most further discussion of false twist
machines in this chapter to that form.
There is a degree of self-adjustment in the angle β. However, under unstable
conditions, there is surging and the angle fluctuates. At high speeds, torque and
tension surges lead to difficulties and impose a limit on the speeds that can be
achieved. A feedback mechanism involving the phase relationships between the tension
and the rotational speed of the yarn leads to the surging.
Equation (4.4) shows that the degree of texturing is strongly affected by the coefficient
of friction, the linear density of the yarn being textured, the applied yarn tension, and
the yarn angles. The angles α and γ may not be the same, but for the purposes of
explanation let them be typified by a single value, θ. The twist level is also a function
100
Handbook of yarn production
of the stiffness of the yarn, as well as the torque. For a given yarn, it is important to
use high values of µ and θ. To give high values of µ, bushes or disk tires made of
urethane or some other high friction material are used. It is difficult to get a high
value of θ with a single disk (<90°) and stacked disk twisters such as those shown in
the center of Fig. 4.4(b) are usually used to give high cumulative values of θ. With a
simple bush, θ is limited and the amount of relative rubbing at the bush ends becomes
a problem because the rubbing causes accelerated wear. The tensions must also be
limited, otherwise there is likely to be individual filament breakage caused by the
excessive friction.
Considering the stacked disk type of false twister, the outside surfaces of the disks
are the drive surfaces. A high cumulative value of torque is obtained as the yarn
follows a sinuous path through the stack of disks. The multiplicity of disks makes it
possible to generate sufficient torque in the yarn to produce the desired texture in the
material. But there can be a progressive increase in yarn tension, which (if allowed to
get too high) can cause damage to both the yarn and drive rollers. Generally, the
stacked disk type of machine can operate commercially between about 15 denier (17
dtex) and 150 denier (167 dtex), at threadline speeds (V) of the order of 500 m/min.
As was pointed out, the angles of the threadlines are important. The cumulative
value of θ in a stacked disk arrangement is dependent on the depth of penetration of
the disks. The angle β is also affected. Some designs use three sets of disks with
equidistant centers; the distance apart of the sets of disks (i.e. the penetration) is
adjusted by using various spacing bushes. Another design has one set of disks hinged
so that penetration can be easily adjusted without having to ‘re-string’ the system.
The hinged stack system also makes stringing up much easier because one set of
disks can be swung out of the way to allow insertion of the yarn. Some designs use
a number of smooth surface guiding disks that serve to merely guide the yarn through
the stack. These disks are adjusted to give the desired run-on and run-off angles at the
working disks (i.e. the angles α and γ). The guiding disks supply little or no torque
to the yarn. A variety of disk profiles can be used, and the driving disks have a variety
of drive surfaces.
Because of the relatively high cumulative values of θ, it is possible to replace the
rubber-like surfaces with a more durable, hard surface. The most successful of these
hard surfaces to date seems to be aluminum oxide (Al2O3) but other possibilities
include plasma coatings, various other oxides, glass, glass mixtures, ceramics, synthetic
rubbers, and polyurethane. Also under development is the use of artificial diamond
dust embedded in nickel. Always, the balance to be considered is between the coefficient
of friction obtainable and the wear rates of both disks and yarn.
If the angle of the disks is changed so that a component of the frictional forces acts
along the threadline, the disks tend to pump the yarn through the system without
large increases in tension. Also, if the yarn can be encouraged to work at an angle β
(Fig. 4.7(b)), a similar result is obtained. In practice, the yarn lies at the angle β quite
naturally, and the value is affected by the disk penetration. Thus, there can be a degree
of pumping even with parallel disks, and so most practical disk texturing systems are
carefully designed to allow the yarn to pass through with moderate tensions. The
accumulation of the angles of wrap through the stack causes the torque available to
the yarn to increase without a corresponding increase in tension. Limiting the yarn
tension improves efficiency, decreases filament breakage and reduces wear of the
disk driving surfaces. However, if the yarn tension is allowed to drop too low, there
can be a loss of control, which causes problems. The normal tension ratio between
Filament yarn production
101
input to the disk stack and output is 1.5. Also, the torque produced tends to drop over
time, as the surfaces become worn and slick.
Another factor that requires special vigilance is the change in frictional characteristics
of the disks. As the surfaces wear or become polluted with polymer or breakdown
product, the frictional characteristics change. A good drive surface tends to wear
clean but there is still a tendency for changes to occur even though they happen much
more slowly. Soft surfaces, like polyurethane, can easily be damaged by inexpert
handling; also a wrong setting causes very rapid deterioration. The hard surfaces are
more durable and the damage is much more likely to occur to the yarn. In particular,
filament breakages can be very troublesome. Variations in the torque can vary the
hand and appearance of the fabrics made from the yarns.
As was mentioned earlier, there is some slip in friction twisting and the exact
amount depends on the cumulative values of µ, β, θ, and T, as well as on the operating
speed. Since µ and T are limited, the major variables are the operating speed and
depth of disk penetration (which affect θ and β). Although variations in µ due to
changes in humidity or fiber finish might be considered to be minor when compared
with those of speed and penetration, they cannot be ignored because they directly
affect the quality of the product. The fiber finish can be heavily modified, or even
burned off, by overheating. In terms of quality control (rather than machine design),
variation in µ is important. Some effects of variations in θ and speed are given in Fig.
4.8. For very high production rates, V must be high and the practical variable becomes
θ. Too high a value of θ causes high end-breakage rates and unsatisfactory yarn,
which is why the length of the bottom curve is so short. There is exceptionally high
filamentation (i.e. breakage of filaments, where many of the filament ends appear as
hairs on the surface of the yarn) under the latter conditions mentioned. Consequently
there are upper limits to speed and torque. Productivity is very high but it is limited,
despite the fact that the twist is applied directly to the yarn surface. The slip is
roughly an exponential function of the twist density (tpi or twist/m); at high twist
θ = high, V = low
θ = low, V = low
Torque
θ = low, V = high
θ = high, V = high
Theoretical twist insertion rate
In-service time of texturing element
Fig. 4.8
Torque produced by a stacked disk twister
102
Handbook of yarn production
levels, the slip level may approach 50%. This not only causes wear but also raises the
temperature of the yarn to dangerous levels.
One design variation is to use a grooved ball that meshes with one or more disks.
The yarn torque accumulates in much the same way as already described. However,
the larger surface area on the ball distributes the wear and allows the higher coefficient
of friction associated with a softer material to be used. Another variant uses crossed
belts to apply the twist. These keep good control of the filaments but belt wear can
be a problem.
The input and output velocities Vin and Vout in Fig. 4.7 differ because of contraction
and the feed and the take-up have to be adjusted to take this into account.
4.5
Draw-texturing
As texturing speeds rise, they approach the speeds used for filament drawing and it
becomes possible to contemplate a merger between the two operations. This raises
the question of whether the fiber producer or the throwster should do the whole
operation. (The throwster is a person or organization that carries out only the texturing
operation. It was derived from the silk trade.) It may be recalled that the freshly
extruded filament is relatively weak and ages rapidly. However, at high extrusion
speeds, the polymer does become partially oriented and filaments might be stable
enough to ship to the throwster. If the feed yarn is partially oriented (draw ratio ≈
1.7), ageing is a relatively minor problem. The use of partially oriented yarn (POY)
as a feedstock for the throwster is quite practical provided proper care is exercised in
inventory control and it is now a firmly established procedure. Databases are often
used to ensure that the material is used in timely fashion and that none of the feed
yarns remain after their shelf-life has expired. Once such logistical problems are
solved, there are several benefits to the use of POY, as was discussed earlier.
There are two forms of draw-texturing; namely, (a) sequential, and (b) simultaneous.
In the former, the drawing and texturing are separate phases within the same machine,
whereas with simultaneous draw-texturing the drawing, heating, and twisting are
carried out simultaneously (see Fig. 4.9). Simultaneous draw-texturing may be carried
out on a conventional texturing machine by merely altering the feed and take-up roll
speeds. Although it is cheaper to use simultaneous draw-texturing, the yarns are
drawn in the twisted hot state in this process, which results in a variation in the draw
from one filament to the next. There can be an inferior degree of setting and a poorer
crimp-resilience; also, at high speeds it is difficult to get a sufficient draw in all
filaments without excessive tension. As has been discussed, the high tensions cause
filament breakage or even end-breakage and this not only impairs the quality of the
product but also impairs the operating efficiency. However, the economics of the
situation favor simultaneous draw-texturing.
One problem is due to flats that develop on the filaments and give the yarn a
crisper hand than a pin twisted yarn, and a different optical effect. It is claimed that
draw-textured yarns are less prone to barré and the picking, pilling, and snagging
associated with knit goods, provided that there is good control over the age of the
feeder yarn. It is also claimed that higher bulk can be achieved, and that more level
and deeper shades of dyeing are possible. It will be noted that there are pros and cons,
but the balance has swung in favor of friction twisting and draw-texturing. The
combination has become an important texturing system. There are variations on the
Filament yarn production
103
120 m/min
120 m/min
Godet
Draw
Godets
400 m/min
Heaters
Draw
False
twisters
Godets
400 m/min
400 m/min
Heaters
Godets
380 m/min
380 m/min
Winders
POY
POY
(a) Sequential
(b) Simultaneous
Fig. 4.9
Two forms of draw-texturing
theme, and very likely there will be more, but this book can only deal with the
principle. However, it is interesting to note that a number of draw-texturing systems
have run commercially above 450 m/min for some time, and this is equivalent to
0.4 lb/spindle hr (0.18 kg/spindle hr) when producing a 70 denier yarn. Speeds of
over 1000 m/min have been reached in the laboratory.
At very high speeds, there can be surges of twist and tension, which adversely
affect the quality of the yarn. Careful control of all the parameters is necessary to
avoid these instabilities. Also, disturbances, such as knots, can provoke instability
and there may be a considerable amount of faulty yarn processed following the
passage of a damaged section of yarn, knot, etc. For this to happen, the machine has
to be operating near the critical range of speeds, tensions, and twists. The higher the
speed, the more difficult it is to avoid the problem.
104
Handbook of yarn production
4.6
Stuffer box texturing
4.6.1 Fiber buckling
The stuffer box has long been used to texture yarns and fibers, but modern technology
has caused it to again become interesting outside its original usage. For this reason,
it is desirable to explain the underlying principles. In essence, a yarn is overfed into
a heated chamber and the overfeed causes the hot filaments to buckle. They become
set in that configuration as they cool, perhaps before being removed from the stuffer
box. It will be recognized that the phases of heating deformation and cooling have
again appeared, except that now the deformation is a zigzag type of crimp rather than
coils or snarls.
In this new case, there is no need for twist and extremely high speeds become
possible. However, to maintain quality, the size of the zigs and zags have to be
controlled, otherwise the variance in crimp affects the appearance and hand of the
product. The filaments in the stuffer box just before buckling behave as struts. Figure
4.10 shows a long, slender filament subject to end loading. A small load, F, causes a
deflection, y, which causes a bending moment at the mid-point of the fiber. The
deflection y increases uncontrollably when buckling occurs; ends move together to
produce a crimp. The actual crimp amplitude and the crimp frequency are defined in
the lower portion of Fig. 4.10 and the maximum amplitude is A. The system is
unstable and the strut tries to collapse into parallel portions, each of length l/2. In a
constrained situation, the filament collapses into a zigzag shape as shown in the
bottom portion of the picture. The length, l, depends on the design of the machine and
the size of the filament.
Buckling length for round strut > (Cross-sectional area) × √πE/F
[4.5]
This suggests that the crimp is dependent on three major factors, namely: the buckling
force F, the modulus of elasticity E, and the geometry of the cross-section. The force
F is principally determined by the degree of overfeed. The polymer and its heat
treatment determine the modulus. The geometry of the cross-section is established
during extrusion and is a function of the linear density of the filament. Thus the
texture is seen to depend partly on the feed rates and the temperature within the
stuffer box.
F
y
F
l
A
l/2
l
Fig. 4.10
Fiber collapse in a stuffer box
Filament yarn production
105
4.6.2 Stuffer box
Some modern systems depend on a controlled overfeed and a fiber transport system
within the stuffer box such as is shown in Fig. 4.11. An overfeed is a condition where
the input speed is greater than the speed further along the process flow line. The
transport system is intended to improve the uniformity of the process at high speeds.
Without it there can be a tendency to intermittently choke. Even partial chokes affect
F and thus the crimp level. Hence, a smooth flow of fiber through the stuffer box is
essential. Another difficulty at very high speeds lies in ensuring that each filament is
heated to the same temperature. Not only is it necessary to raise the filaments above
Tg, but all filaments should have identical temperature histories so that conditions are
the same for all. Failure to provide such conditions leads to a variation in crimp level
from filament to filament. Although it is not feasible, in practice, to transfer the heat
equally to all filaments, at least the variation should be kept to a minimum.
Some fine stuffer box textured yarns are plied to give the material a resistance to
snagging and filament breakage in the fabric during normal use. However, plying is
expensive and there is a loss of bulk in the yarn (which was the purpose of texturing
in the first place). Sometimes the bulk is not fully developed until the fabric has been
finished and this means that some potential faults are not discovered until the fabric
finishing process is completed. Omission of a heat setting stage, or the use of improper
Flat yarn input
Vin
Vin > Vout
Stuffer box
feed rolls
B
Tractor feed
transports
fibers
through the
stuffer box
Heat
VS
Cool
A
Controller
Vout
Winder
Fig. 4.11 Stuffer box texturing
106
Handbook of yarn production
processing temperatures, cause changes in bulk and dye affinity, both of which can
lead to barré in the fabric. Thus it is necessary to test the product [3] for shrinkage
and dye affinity at the yarn processing stage to avoid expensive claims from customers
because of improper quality.
It has become possible to process yarns at up to 1200 m/min. This may be compared
with the speeds obtainable for friction twisting. Unfortunately, the crimp stability and
the uniformity of stuffer box yarns is not so good as with false twist textured yarns.
Nevertheless, the system is capable of handling relatively heavy yarns so it has
become quite important in carpet yarn manufacture [4]. A stuffer box takes up little
space and can easily be placed in line with another process. Because of the high
speed capability, it is often used for crimping tow. It would be very difficult to do this
with other methods because of the large number of filaments involved.
Hot fluid texturing is a variant of stuffer box texturing, where the solid filaments
in the stuffer box are replaced by jets of hot fluid polymer. As the material enters the
nozzle in a plastic condition, the strands are looped or otherwise disturbed before
they impinge on the plug of filaments in the stuffer chamber. The outgoing yarn is
wrapped around a cooling drum to set the crimp. This is a form of bulked continuous
filament (BCF) production, which spins and texturizes the filaments in one operation;
it is used mostly to produce nylon and polypropylene yarns for floor coverings [1].
4.7
Air-jet texturing
4.7.1 Simple air-jet devices
All the foregoing methods of texturing require that the yarns be thermoplastic so that
they can be heat set. This precludes the use of non-thermoplastic yarns like rayon.
Air-jet texturing provides a means of creating texture in such materials. Further, it is
a useful means of producing a yarn structure near to that associated with staple yarns.
This is an important concession to the tastes of the ultimate consumer. False twist and
air-jet texturing can be combined.
The major principle involved is the tangling effect given by highly turbulent airflow
acting on filament feed yarns. Entanglements within the yarn structure are made, and
are interlocked by inter-filament friction to form a stable yarn. In some ways, these
air-textured yarns resemble staple yarns made by traditional spinning methods. To
get the needed air turbulence, high pressure air is supplied to a nozzle and this
produces supersonic airflow at the exit. Also, an obstruction or asymmetry is introduced
in the airstream to cause a series of violent eddies; this is known as a von Karman
vortex stream. The obstruction can be in the form of a hollow needle through which
the feedstock is fed. Because the emerging airstream contains shock waves (like
those seen in jet engine exhausts), there are some severe pressure gradients in the air
discharge. A diagram of the divergent portion of a nozzle with a filament injection
needle is shown in Fig. 4.12(a) where the swirling airflow (gray arrows) passes over
an obstruction such as needle, creating turbulence downstream (shown in black). The
attitude of the needle, and its rotational position about its own axis, are adjusted to
maximize the quality of the textured yarn. Because the needle is hollow, it acts as an
injector since the static air pressure in the throat of the nozzle is less than atmospheric
pressure. Thus, a filament feed yarn can easily be inserted into the exit airstream
(Fig. 4.12(b)). Separated filaments follow different flow paths and when the filaments
are recombined at an integration point, there are lengthwise displacements of one
Filament yarn production
107
Textured yarn output
Filaments
separate
from each
other
Needle
Yarn input
Airflow
(a)
Fig. 4.12
Airflow
(b)
Air-jet texturing
filament to another; some filaments are overfed and the result is that a structure with
loops and bows is formed, as shown in Fig. 4.13. A bow in this context means a
curved portion of filament that does not make a complete loop.
The needle causes the airstream, which is passing over it at high speed3 to break
up into eddies. These eddies can be superimposed on a general vortex motion tending
to untwist the feed yarn. The untwisting allows separation. However, separated filaments
possess torque because of the untwisting and, if overfed, the filaments tend to curl or
snarl and occupy more space. Since the filaments are separated, different filaments
are caught by the progression of eddies and there is a tangling effect as the snarls and
loops become caught up in each other. Filament separation is an essential part of the
texturing operation. The subsequent tension applied to the filaments after they recombine
at the integration point causes the loops and tangles to interlock to give a moderately
bulky yarn. The yarn has characteristics similar to staple yarn. Longitudinal migrations
of portions of the filaments, caused by differing path lengths taken by the filaments
between separation and integration, enhance the texture because some filaments are
temporarily overfed with respect to their neighbors (in Fig. 4.14, filament a has been
overfed with respect to b and c.) The excess lengths produce loops and bows. Compared
to false twist textured yarns, air jet yarns are considerably less extensible.
In some operations, the entering filaments are moistened, which enhances the
texturing operation because of better separation of filaments within the nozzle; control
of the flowing filaments is also improved. This is referred to as the wetting process,
where one or more yarns pass through a water bath before entering the air-jet. Care
has to be taken to remove the debris or finish particles that accumulate, so that the jet
nozzles do not become blocked. Alternatively, water applicators are used, which
allow finer control of the water applied.
A baffle is sometimes used to divert the flow, to create extra turbulence and to
3 The Reynolds Number must be above the critical value.
108
Handbook of yarn production
Fig. 4.13
Air-jet texturing yarn
Integration point
Bows
Filament
migration
c
Separation
of filaments
b
a
Nozzle
Fig. 4.14
Filament separation
Filament yarn production
109
lessen air consumption. Baffles can be used to limit the filament bow size and control
the loopiness of the yarn. Bearing in mind that stability of the yarn depends on interfilament friction, it might be realized that a drawing stage following the texturing can
stabilize the structure by pulling the closely looped portions tighter. The drawing
process in this case is like tightening a knot. A thermal process may follow the
texturing [5] to achieve a reduction in loop size and to reduce shrinkage in boiling
water.
4.7.2 Effect yarns
As a class, effect yarns are a speciality of interest to fabric designers looking for
special effects in their products. Yarns with nubs, bouclé yarns with loops on the
surface, and many more, are members of the class. It is beyond the range of this book
to deal with them all, but a few processes will be mentioned in passing to give a flavor
of a few possibilities.
There are special mechanical attachments that can be fitted to normal spinning
machines to produce effects such as aperiodic nubs or loops. Some of these are based
on a random speed varying device that affects the draft in staple spinning. However,
these are not very useful when drawing a filament yarn because of the variation
caused in the molecular structure. More likely one will find devices that raise loops
or break them to produce the desired effects. There are also some treatments based on
unequal shrinkage of components within the yarn structure to produce bulk, perhaps
in a randomly induced fashion. Air-jet texturing is sometimes used in series with the
basic yarn process. Some spin staple fibers to form a sheath around a core of filaments;
these (together with those described later in this section) are called core yarns. Such
core yarns are sometimes regarded as ‘effect yarns’ when they produce special effects
rather than act as replacements for traditional yarns.
If the components within the combination of fibers or filaments can be induced to
shrink differentially with respect to one another, then extra bulk can be produced,
sometimes evenly and sometimes not. If some fibers are capable of being set and
others are not, then a further set of possibilities arise.
Slitting or fibrillating thin polymer sheets may make flat filaments, like miniature
ribbons, which can then be made into yarns. Fibrillation may be carried out by
drawing a sheet of certain polymers such as polypropylene and concurrently applying
lateral stress to produce a yarn of flat filaments without the need for slitting. These
so-called flat filaments may be mixed with some of those already discussed to produce
interesting visual effects arising from their differing optical properties. Combinations
of various of the yarns described in the various sections bring the possibility of a
wide range of effects.
The idea is extended by extruding different polymers through the same spinneret
and combining them as a ‘co-extruded yarn’ (see Section 4.8.6). Alternatively different
spinnerets are used for each polymer and the filaments are mingled together before
taking-up prior to winding to produce a ‘co-mingled yarn’. For example, it is possible
to use a component to give strength in the core and a more aesthetically pleasing fiber
as the sheath. The component delivered to the nozzles at the highest delivery speed
is the ‘effect’ component, which goes mainly to the sheath, and the component fed at
the lower speed becomes the core. (The more slowly moving filaments approaching
a mingling point are under more tension than the faster ones, which produces a
110
Handbook of yarn production
migration similar to that described in Section 3.9.3.) It is possible to use POY as one
of the components and to include a drawing stage in the process.
4.7.3 Modified false twist texturing
Air-jet texturing is now being used in conjunction with false twist texturing to produce
filament yarns with staple-like characteristics [1]. Modifications to the structure
involve surface loop control and/or the production of free fiber ends in the surface to
simulate staple fiber yarns. Feeding two or more sets of filaments into the yarn at
different rates can form loops, and also modifying the polymers can change yarn
properties. The conditions in melt spinning can also be varied to alter the structure.
The ability to extrude very fine filaments has also increased the range of possibilities.
The great number of alternatives not only makes the modern machines much more
complex than formerly but the technology draws on a much wider base. The result is
a wide range of product possibilities. Control of fiber speeds, tensions, and temperatures
at all positions is an essential prerequisite for consistent and acceptable yarn quality.
To get high productivity and adequate bulk, it is necessary to use expensive high
pressure air. Also, to control bulk, it is essential to maintain the settings, which uses
expensive labor. On the other hand, the air texturing produces no appreciable
morphological changes in the polymer and at least one source of barré is removed.
Productivity is very high.
4.8
Other texturing techniques
4.8.1 Bi-component yarns
The basic idea of a bi-component yarn is to use filaments that consist of two parallel
components, each having different physical attributes (which affect their shrinking or
swelling characteristics). A composite structure has the potential to curl if a filament
consists of polymers A and B disposed side by side as shown in Fig. 4.15(a). The
filament curls when polymer A is caused to shrink relative to polymer B. This is
because of the forces generated by the shear due to shrinkage. If the differential in
shrinkage is sufficient, and the ends of the filament are restrained, the curl develops
into the reversing-coil helix sketched in Fig. 4.15(b). As with other textured yarns,
this improves the bulk and lowers the effective modulus of the yarn. However, the
result is obtained without mechanical texturing and therefore is not restricted in the
same way. There is potential for very high speed production, but the method is often
applicable only to very fine yarns.
One method of producing such a structure is to extrude compatible but different
polymers through the same spinneret. It is important that the components mutually
adhere. This rules out using polyester at the present. Usually two forms of nylon are
(a)
(b)
Fig. 4.15
Bi-component yarn
Filament yarn production
111
used. Another method is to combine two dissimilar strands from adjacent spinnerets
in such a way that they adhere to produce a bi-component yarn. Again, it is very
important to make certain that there is adequate bonding between the components. A
considerable volume of such bi-component yarn is used for ladies’ hosiery.
4.8.2 Edge-crimping
A product related to bi-component yarn, but not always regarded as such, is edgecrimped yarn. If a yarn under tension is run over an edge (Fig. 4.16), a lengthwise
layer of polymer is disoriented and possesses different shrinkage characteristics from
the rest of the yarn. The effect can be demonstrated by running a human hair over a
finger nail and watching it curl. One of the problems with an edge-crimping process
is the maintenance of the edge over which the yarn slides. Variations in conditions at
the edge lead to variations in crimp and thus to quality control problems.
A further related idea is that of asymmetric quenching of the yarns at extrusion (or
elsewhere). The rate of cooling affects the crystallinity and is associated with variations
in density. In other words, asymmetric quenching can also produce a texturing effect.
It is believed that similar effects could be produced chemically. In any of these cases,
the bulk can be developed by heating, which can cause further differential shrinkage
(or swelling) to augment the effect.
4.8.3 Twisting and folding of filament yarns
It should be explained that ‘folding’ in this context is jargon used in the filament
trade; it has a similar meaning to the ‘doubling’ discussed earlier, inasmuch as strands
are laid more or less side by side before they are integrated into the final yarn. The
process is often a two-step operation with a forming twist being first applied to single
ends and then cable twisting the composite to achieve the desired end result. The
final product has a low or zero filament twist, but the ply twist is sufficient to control
the surface of the yarn. Often two-for-one twisting or a variant of it is used for these
operations. There is little or no need for the improvement in evenness that such
doubling brings. Reasons for this operation include [6, 7]: (a) entrapment of wild
fibers or broken filaments, (b) torque balancing of false twisted yarns, (c) improvement
of load sharing between the filaments, (d) changing the load elongation characteristic
of the yarn, and (e) changing the optical and tactile character of the yarn.
Edge
Filament
yarn
input
Oiler
Textured yarn delivered
Fig. 4.16
Edge-crimp texturing
112
Handbook of yarn production
4.8.4 Knit-de-knit texturing
The fundamental idea of knit-de-knit texturing is simple. If a fabric is knitted, heated,
and cooled and thermoplastic yarn is unraveled from the fabric structure, then the
yarn is found to have a texture set into it. The newly unraveled yarn has repeating
deformations, but these can be manipulated to redistribute the zigs and zags of
individual filaments and create a textured yarn. It is used for certain specialty yarns.
For example, where low bulk, lustrous fabrics are required using a fiber such as
Quiana® (a high cost nylon used as a high fashion silk substitute), then the knit-deknit process might be appropriate. In such specialty markets, it is aesthetic results
that are more important than high productivity and low price.
4.8.5 Elastomeric yarns
Elastomeric fibers are characterized by very high elongations at break (up to 100%)
and have a composition of at least 85% segmented polyurethane [8]. They owe their
extensibility to the soft, elastic material used. Polyethers or polyesters are used as
segments of block co-polymer chains, which are joined together by urethane groups
but which are not cross-linked. The result is a polymeric structure capable of high
‘power’ yet which can be heat set into desired shapes. In this context, ‘power’ refers
to the ability of the material to recover from elongation or other deformation. A large
proportion of this material is used in foundation garments, swimwear, and hosiery.
Sometimes an elastomeric core is sheathed with another type of fiber to give good
aesthetic properties. Care has to be taken that the elastomeric core does not ‘grin’
through to give unsightly changes in color or reflection due to different dye behaviors.
4.8.6 Texturing by co-extrusion
Co-extrusion is where two or more polymer components are extruded through the
same nozzle to produce a filament with stripes of different polymers (Fig. 4.17). It is
difficult to manage more than two components; thus two component systems are
likely to be most significant commercially. There are two distinct possibilities. The
first is to have the stripes firmly bonded to each other in such a fashion that treatment
will cause it to curl or otherwise texture in the manner of a bi-component yarn. The
second is to make the stripes have little or no bonding, in which case the filament can
be decomposed into a series of finer ones. Ultra fine filaments can be separated from
the main body to make silky yarns and a variety of surface effects are possible by
altering the cross-sections of the separated fibrils. Multi-lobed cross-sections diffuse
(a)
(b)
(c)
Fig. 4.17 Co-extruded filament yarn and components
Filament yarn production
113
reflected and refracted light to give a dull effect whereas flat cross-sections give a
sparkle such as that associated with silk. The author has no details of the production
of these materials.
4.9
Industrial filaments
Polypropylene (an olefin) is sometimes used for some non-apparel yarns but care has
to be taken to protect the yarns from sunlight, which degrades them. The moisture
absorbency is less than 1%, which is a serious disadvantage for apparel and some
home uses. However, it does have good dimensional stability if the temperature is
kept below about 120°C (≈ 250°F). The main use is in industrial fabrics. For that
reasons there is little need to consider texturing the yarns.
High tensile man-made filaments, such as those made from aramid polymers, are
also used for many industrial applications, such as ropes and cables, because of their
very high tenacities. Other common industrial filaments are those of polypropylene
and similar polymers, which are used for carpet backings, bale wrappings, etc. Space
precludes discussion of the technical aspects of ropes and cordage but the reader is
referred to the work of Backer [9].
Other fibers are used because of their modest cost and/or their high strength. Glass
and high modulus, high strength fibers, such as carbon, are increasingly used for
reinforcement of composites but discussion of this important sector must be curtailed
because it carries us beyond the production of yarn. When sheets of certain polymers
are stretched, they split in the direction of stretch with a result that the sheet is
transformed to a web of interconnected filaments. This process is called fibrillation
and it was discussed briefly in Section 4.7.2. The use of chopped fibrillated material
falls outside the range of our discussions although some fibrillated materials do end
up as yarn, even if only in tape form.
Often these fibrillated filaments have a rectangular cross-section. Sometimes the
position of the slits is precipitated by ridged roll surfaces, or the sheets are slit.
According to Schuur and Gouw [10], it is a pity that water bath quenching is less
suitable for making thin films because of draw resonance, which gives unacceptable
thickness variations. In other words, it seems that it is not yet possible to make fine
fibrillated filaments. The stretching of the film is carried out in ovens with forced-fed
hot air. A stretching force of 1 to 2 g/den (i.e. 9 to 18 g/tex) is normally used.
Sometimes bi-component structures are created by using laminated sheets of different
polymers, e.g. polypropylene and polyethylene. This gives a structure that is easily
textured to give bulk. If the sheets are slit into narrow strips, the result is a textured
yarn. Untextured strips of polypropylene are used directly as yarns where more
robust use is contemplated, as in the manufacture of sacking, bale coverings, carpet
backings, and the like.
4.10 Silk filaments and staple yarns
Silk filaments are converted into yarn by a process known as throwing.4 The filaments
from the skeins arriving from reeling in the filature have to be plied. This requires a
4 From the Anglo-Saxon ‘thrawan’, to twist.
114
Handbook of yarn production
twist of perhaps 4 or 5 tpi (0.1 t/m) to be added during the plying process. The plied
yarns are then twisted to the level required for the end use. Twisting is sometimes
carried out by ring frames similar to that shown in Fig. 7.3, but sometimes there is a
twister included in the reeling equipment that produces hanks of silk yarn. In many
of the silk producing areas of the world, silk goods are an encouraged cottage industry.
In those areas, there is still a considerable amount of manual manipulation of silk
filaments in the production of yarn. Staple yarns are often thrown using spinning
wheels and mule spinning frames.
The plied silk yarn usually has considerable amounts of gum left on it, and it is
quite normal to produce a warp yarn that needs no sizing for weaving. Most other
staple yarns and some filament yarns need to be sized by the addition of a softened
adhesive to withstand the rigors of weaving.
4.11
Morphology and dyeing
Dyes are color producing substances that can be permanently attached to or incorporated
into the fiber. The affinity between the dye and the fiber depends on the physical and
chemical properties of both. As has already been mentioned, the physical characteristics
of the fiber depend upon its mechanical and thermal history. The morphology of a
polymer changes as it is heated and cooled. It also changes as the fiber is drawn. The
dye affinity of the material changes accordingly. Thus, the texturing operation can
affect the dyeing operation materially. If there are periodic variations in polymer
morphology arising from any of the manufacturing stages preceding the dyeing
operation, there will be periodic changes in the color of the yarn along its length. If
the wavelength of the error is small, the fault appears in the fabric as a moiré effect.
If the wavelength of the error is large, the fault appears as barré. Such periodic errors
could be caused by finish deposits on a feed or take-up roll in the texturing, or by
faulty winding, or some other mechanical error. Many yarns are dyed in the form of
relatively low density cones or cheeses and the winder on the texturing machine has
to be configured accordingly. Staple yarns are sometimes dyed in hank form. Thus,
if there is uneven dye penetration into the package, a range of error wavelengths may
be found from this cause also. It is possible, and desirable, to determine these wavelengths
by dyeing a knitted test sleeve, or by other means, to find the source of the problem.
In addition, there can be more random types of variation arising from a variety of
causes, such as spindle-to-spindle variations in the texturing conditions, mechanical
or thermal instabilities in the texturing machines, faulty winding, etc. These variations
tend to show up in the fabric as shading or streakiness.
References
1.
2.
3.
Wilson, D K and Kollu, T. The Production of Textured Yarns by Methods other than the Falsetwist Technique, Text Prog, 16, 3, 1987.
Atkinson, C and Wheeler, M J. New Developments in Air-jet Textured Yarns for Upholstery,
Int Text Bull, 1, 1996.
Du, G W and Hearle, J W S. Threadline Instability in the False-twist Texturing Process, J Text
Inst. 81, 1, 36–47, 1990.
Filament yarn production
4.
5.
6.
7.
8.
9.
10.
115
McCormick, W H. Bulked Yarns Produced by a Stuffer Box Method, Modern Yarn Production,
(Ed G R Wray), Columbine Press, Buxton, 1969.
Bock, G and Lünenschloss, J. An Analysis of the Mechanisms of Air-jet Texturing, Textile
Machinery: Investing for the Future, Textile Inst Ann Conf, 1982.
Fischer, K E and Wilson, D K. Air-jet Texturing – An Alternative to Spun Yarn Production,
Textile Machinery: Investing for the Future, Textile Inst Ann Conf, 1982.
Lorenz, R R C. Yarn Twisting, Text Prog, 16, 1/2, 1987.
Craig, R A and Ibrahim, S H. Elastomeric Fibers, 4th Shirley Int Seminar, The Hague,
Netherlands, 1971.
Backer, S. The Mechanics of Bent Yarns, Text Res J, pp 668–81 and a number of later papers,
1952.
Schuur, G and Gouw, L H, Future Prospects for Fibrillated Polypropylene Film Processes
and Products, 4th Shirley Int Seminar, The Hague, Netherlands, 1971.
5
Carding and prior processes for
short-staple fibers
5.1
Introduction
Short-staple fibers are nearly always processed dry using mechanical means. The
first stages in a short-staple spinning mill comprise a number of machines, usually
arranged in series, which are connected by fiber transport systems. The most common
of these transport systems is where air is pumped through large ducts and carries the
fibers in the airstream. The line of machines described is called an ‘opening line’ and
it supplies a set of cards in which the fiber flows are usually arranged in parallel.
There are usually two or more opening lines in an establishment and the space
occupied by opening lines is sometimes known as the ‘blow room’.
One function of the blow room is to blend the fibers into a homogeneous product.
The term ‘blend’ applies to the mixing of nominally similar fibers or to the mixing
of unlike fibers similar to polyester staple fibers and cotton. In the latter case, the
blending may be carried out in the blow room or in a process following carding.
However, the case being considered in this chapter is that of blending nominally
similar fibers. It is possible to blend dissimilar fibers by using the same process as
described in this chapter although many operators prefer to do it in processes following
carding as will be described in Chapter 6.
Fiber attributes vary from bale to bale and within a bale. Even man-made fibers,
in which the fiber length and fineness are strictly controlled, have variations in fiber
crimp and finish. Variations of crimp and finish alter the mutual cohesion of the
fibers within a strand or clump and these can strongly affect the ease of processing.
Natural fibers vary in all of their attributes. In both cases it is very desirable to blend
as early as possible in the blow room, and to use every succeeding opportunity to
carry on the process of blending.
The primary stage of blending is carried out by removing clumps of fibers from a
succession of bales and mixing them in the machines that follow (see Section 5.4).
The secondary stage is carried out in a blending machine with the intention of
homogenizing the material in transit. Mixing also occurs in every machine in the
opening line as well as during transit. Further blending occurs in processes following
Carding and prior processes for short-staple fibers
117
carding as will be seen later in Section 6.4 in Chapter 6. All contribute to the degree
of fiber homogenization in the total spinning process; however, for now we must
concentrate on the blow room.
Raw material is supplied to a mill in highly compressed bales of fiber. One important
function of the blow room is to disintegrate these bales into a flow of very small
clumps of fiber, which are sufficiently small in size to be digested by the cards. The
cards then further divide the clumps into single fibers (or very small groups of them)
and assemble them into rope-like strands called ‘slivers’. The function of breaking up
the bales into clumps, and the subsequent reduction into single fibers, or very small
groups of them, is referred to as ‘opening’. (Also the term is used as an alternative
to ‘blow room’ but the context usually makes it clear which meaning is intended.)
A third function is needed for natural fibers, the most important of which in shortstaple spinning is cotton. As was discussed in Section 2.2.1 in Chapter 2, cotton
ginning is imperfect as far as removal of the trash is concerned. Ideally, unwanted
matter must be removed so that it neither interferes with operations nor causes significant
deterioration in the quality of the product. In practice the ideal is not reached but
modern technology allows a close approximation to it. The mechanical cleaning
function is not required when spinning 100% synthetic fibers, but many mills have
this broad capability irrespective of the fiber actually being spun; this gives them
operational flexibility.
Recombination of fibers into a larger mass occurs at various stages along the
opening line and I will call this phase ‘condensation’. Such condensation is necessary
to accommodate the control of the fiber flow in a continuous line and to aid the
processes of accumulating fibers to make the feed systems workable. Feed systems
often use moving lattice aprons to collect fibers deposited from streams of air or
gravity feeds. A lattice apron is an endless permeable belt of slats, each of which is
positioned perpendicular to the line of the belt movement, but parallel to one another.
Air is often sucked through the gaps between the slats. The slats contain metal spikes
to retain the fibers. A rotating condenser is a perforated cylinder, to which suction is
applied so as to collect fibers from an airstream. The process of condensation is really
a form of doubling (see Section 3.10.1), which improves evenness along the fiber
stream. The word ‘stream’ is meant to include airborne fibers flowing in ducts, thick
blankets of fiber (called batts or fleeces) being carried by mechanical transport
mechanisms (such as lattice aprons), and sliver being delivered from the cards.
It is impossible to blend the fibers into an intimate blend without opening them
first. A perfect intimate blend would have a single fiber of one sort in very close
proximity to single fibers from each of the other sorts. Imagine trying to blend
clumps of fibers of, say, 1 cu ft in size into a homogenous product. It would be rather
lumpy and the blend would hardly be characterized by the word ‘intimate’! Also,
adequate cleaning of natural fibers is not possible without opening the clumps first.
In the case of cotton fibers, it is relatively easy to remove the trash and dust from the
outside of a clump but it is much more difficult to remove spiky trash or even dust
from inside without damaging the fiber. With so-called cleaning machines, there is,
of necessity, a great deal of opening and a certain degree of blending. Dust and trash
has to be removed from the bale plucking machines. Dust and trash is ejected from
blending machines, the main job of which is to accumulate fibers in reservoirs to
facilitate blending, as discussed in Section 5.4. Dust and trash is removed from socalled opening machines as the clumps are divided. Consequently, each machine in
the opening line performs functions of blending, opening, and cleaning in varying
118
Handbook of yarn production
proportions. Thus the processes called opening, cleaning, and blending should be
regarded as various phases within the operation of each machine and this is distinct
from the labels applied to the machines. However, a major point of this paragraph is
to emphasize that the labels applied to individual machines in a blow room describe
their major function; all the machines perform all three functions but to varying
degrees.
Cards are included in this analysis because they share many of the functions of the
preceding machines, and they are physically part of the modern linked system for
performing the processes described. The whole system described in this chapter is
integrated. Bales of fiber form the input and sliver is the output. In most modern
systems, the fiber is untouched by human hand from the time that the bale is placed
into position in the bale laydown until the sliver emerges from the card. It is now a
continuous operation and there are few demands on labor. Such a system is known as
a chute feed system because the fiber is fed to the card by way of a ‘chute’. (A chute
is really a temporary storage chamber that contains automatic flow control devices to
maintain the linear density.) Any failure to control at this point would eventually
result in exceptionally long-term errors in the yarn. Errors in the yarn arising from
this cause are so long that they extend over the mass of yarn on many consecutive
bobbins. Once the bobbins become mixed with others, the error appears as a random
count variation. Consequently, flow and control will also be discussed later in Section
5.9.
This chapter will be written using imperial units common in the USA and many
other English-speaking areas. However, metric conversions will be given but in a
paragraph with several such conversions, they will gathered at the end of the paragraph
to minimize distraction to the reader.
5.2
Opening line
5.2.1 The elements of the chute feed system
The elements of a system are shown in Fig. 5.1. The diagram is deliberately incomplete
because the space in the diagram was at a premium and the intention was to give an
Opening and cleaning machines
Bale plucker
Bale laydown
Mixer
Cards
Card slivers
Fig. 5.1
Elements of an opening and cleaning line
Carding and prior processes for short-staple fibers
119
impression rather than a prescription. The actual machines installed are determined
as a matter of operational need and preference. For example, the blending machine is
shown as being the last in line before the cards on the basis that the best blending is
achieved when the fiber clump size is very small. However, some prefer to install it
earlier on the basis that good blending aids the processes of opening and cleaning.
Some use more than one blending machine per opening line but the extra cost of the
machines is sometimes hard to justify. Only four cards are shown in the diagram but
in an actual plant there are several times this number; the actual number is determined
by the relative productivities of the cards and the bale plucker (the device that removes
clumps of fibers from the lines of bales). There are usually two or more opening
lines, because this permits a shutdown of one for maintenance, adjustment, or other
purpose without closing the whole mill. In similar vein, the ductwork for the machines
has not been joined up to emphasize (1) that a variety of ductwork transition pieces
are needed to complete the fiber flow circuits and (2) that bypasses are often fitted
which require flap valves and forked ductwork.
Safety is a special concern in the work zones about to be discussed; consequently
a special Section (5.12) about such matters has been added to the chapter.
5.2.2 The historical perspective
Szaloki [1] points out that there were few changes in the design of opening and
cleaning equipment in the first half of the twentieth century. There was then a surge
in development spurred to some extent by the increased need to clean the cotton as
it increasingly became picked by mechanical means. He gives a review of opening
and cleaning equipment as of 1976. Remarkable progress was made during the last
century in developing means of connecting discrete machines into continuous production
systems. A good example is the blow room just described. At the beginning of the
century, it required many workers to control and transfer material from machine to
machine in the series.
5.2.3 Conservation of flow
Opening and cleaning machines have to be connected in such a way that matches the
productivities of the various components. Since the machines in an opening line are
all connected, mass flow has to be conserved. The conservation includes not only the
fibers flowing into and out of any element, but also the trash and dust removed. In
other words, what goes in should come out! Appropriate fiber transport systems have
to be provided so that a continuity of fiber flow and control can be maintained. Also
there has to be a distribution system that connects a series of cards in parallel to the
supply system. The change from a series path to a set of parallel paths is needed
because the equipment in the series path has a much larger production capability than
the individual cards in the parallel paths.
5.3
Bale preparation
5.3.1 Selection of bales from the warehouse
An intimate blend starts with the selection of an appropriate number of bales from
large lots in the warehouse. Lots are usually segregated to provide compatible content,
120
Handbook of yarn production
and a bale withdrawn from a lot should have, in theory, similar attributes to the rest
of the bales in the lot. In fact, there are large variations and part of the art of blending
is to arrange the assignment of bales to the various lots in a way that minimizes the
variance within the lot. The bales withdrawn are arranged in sets to make a laydown
as described in Section 5.4.
5.3.2 Opening (the process of removing straps and bale covers)
Bales arrive at a mill in a compressed state and they are protected by a covering and
stored in a warehouse. The bales are moved to the work area a day or so before they
are needed, the straps and the coverings are then carefully removed, and the bales are
allowed to condition. The bales are said to be ‘opened’. (It will be noted that the term
‘opened’ takes on a different meaning from that defined earlier and care has to be
taken to make sure of the context of the word.) Removal of the straps can be dangerous
if not carried out with proper equipment because when the straps are cut, they relax
violently and injury could result if due care is not taken. Failure to completely remove
the bale coverings can lead to the inclusions of ‘foreign fibers’ that produce faults in
the yarns and fabrics. (Removal of the last vestiges of the covering from the underside
of a 500 lb (≈ 227 kg) bale is not easy.)
5.3.3 Bale conditioning
The conditioning just mentioned allows the moisture content and temperature of the
fiber to approach stability. The bales, freed from the restraints of the straps and
bindings, expand and they are said to ‘bloom’. Bale blooming is a natural process in
which the bale grows in size as the stresses in the fiber, arising from compression,
release themselves. Conditioning not only allows the moisture content of the bales to
approach equilibrium but, in cold weather, eliminates condensation of moisture on
the cold fibers. Damp fibers are difficult to process. At the other extreme, excessively
dry fibers are subject to being damaged.
The storage environment should be at about 70°F (21°C) and the rh at roughly
45%. Care should be taken to avoid storage of cotton in freezing conditions otherwise
the strength of the fibers will be reduced permanently. A conditioning curve for a
typical bale is shown in Fig. 5.2, which shows how the density of a typical bale might
take several days to reach a stable state after the straps have been released.
Bale density *
(arbitrary scale)
3
2
1
0
0
5
10
Days after release
15
N.B.
* Bale density was measured by penetrometer
Fig. 5.2
Bale density changes upon releasing the straps
Carding and prior processes for short-staple fibers
5.4
121
The first stage of blending and opening
5.4.1 The bale laydown
Two or more parallel rows of conditioned bales are laid on the floor in proximity and
this is called a ‘bale laydown’ as was sketched in Fig. 5.1. The overall purpose of
opening is to break down the supply material into an open mass of very small clumps
of fiber that can be handled at carding. The word ‘opening’ is used here in a different
sense from that described in Section 5.3.2. The first step is to remove clumps of fiber
from the bales and this is referred to as ‘bale milling’ or ‘bale plucking’. A bale is a
tightly packed mass of staple fiber usually weighing about 500 lb (≈ 227 kg). Several
packing densities are used to yield so-called flat, standard, and high density bales.
Mills within the USA use standard bales but high density bales are used for transoceanic
transport to conserve space and cost. Flat bales are of low density and are used by
mills close to a gin. It is desirable to allow the bales to bloom sufficiently and to
choose bales of the same relaxed size. Otherwise the first cuts from the bale laydown
will not be according to plan.
Bales are supplied from the warehouse in carefully selected sets designed to minimize
variations in the fiber attributes. To these bales, others containing recycled fiber are
often added. However, the number of these should be strictly limited if quality is to
be preserved. Each bale should be inspected for fragments of wrappers or wire before
they are assembled into a laydown.
Bales should be assembled so that they are in close proximity to their neighbors in
the laydown and similarly oriented to create a compact mass of fiber suitable for the
bale plucker operation. Care should be taken to keep the height of the bales similar.
5.4.2 The bale plucker
The first mechanical processing stage commonly used in the mill today is a patrolling
‘bale plucker’ or ‘bale milling machine.’ Sets of rotating spikes or teeth are used,
which cut into the operational surfaces of the laydown in a manner similar to that
shown in Fig. 5.3(a). A typical cutter and associated press rolls are shown in Fig.
5.3(b). The press rolls are to keep the bale surface firm at the time of cutting. Most
(b)
(a)
Fiber
Cutting head
Bales
Fig. 5.3 Milling a bale laydown
122
Handbook of yarn production
machines use at least one pair of cutters. Often the other one in a pair is a mirror
image and rotates in the opposite direction. Some machines cause one of the cutters
to lift in order to balance the offtake of fiber between the cutters. The changes occur
at each reversal of the bale cutter head (at the ends of the bale laydown). Some
makers stop the trailing cutter because the teeth are facing the wrong way. The depth
of cut is an important parameter in determining productivity and degree of fiber
separation. A setting more than, say, 0.2 inches (5 mm) might cause the top surface
of the laydown to become roughened with fiber tags. These tags are torn off in the
next pass to form unacceptably large clumps of fiber in the offtake. The theoretical
minimum number of bales in a laydown is determined by the adequacy of the blending
equipment and the diversity of the fiber characteristics from bale to bale. Currently
the maximum set by machine design is about 50 bales but improvements in the
technology of bale milling and blending will increase that value. In practice, the
number of bales is set to give a work schedule that is suitable for management of the
personnel and minimizes costs. The greater the number of bales in the laydown, the
fewer the number of laydown changes and the lower the cost of operation. Also, the
use of bale pluckers that deliver fibers in small clumps contributes greatly to the
solution of the problem of opening. The clumps at this stage should be the largest in
the system and the spikes or teeth that temporarily grip the clumps have to be
proportionate in size.
It is desirable that bales in each laydown should be of similar size and density. If
they are not of the same height, the first cuts will differ in composition from later
ones. If they are of different density, the blend make-up in the output will not always
be as predicted from the initial bale data.
5.5
The process of disintegration of fiber clumps
5.5.1 Opening (the process of division of fiber clumps)
In the stages of the opening following the bale plucker, machines with an opening
function have the task of separating clumps of fiber into smaller ones. The sizes of
the clumps, and of the teeth that deal with them, are progressively reduced. In general
terms, grasping clumps of fibers with sets of teeth and dragging the clumps across
another set of teeth or grids perform the opening function. The engagement of a
clump with two sets of teeth in which there is relative movement applies a shearing
action that pulls the clump apart. Since this process alone cannot be seen in any
machine, and the design of the machine is affected by the other processes it has to
perform, further discussion of the machine design will be deferred.
Most machines have feed rolls and toothed elements that are significant drafting
systems. Fibers are caused to slide over one another as fiber clumps are divided and
the resulting daughter clumps or single fibers are removed. The divided fiber clumps,
fibers, and non-lint material are carried to the next machine by the airflow and the
material is discharged into a receiver that might be a chute feed, condenser, or the
like. The receiver has a doubling function as will be discussed in the next section.
5.5.2 Specific volume of the fiber stream
The specific volume of a bale varies. However, the changes are dependent on the
fiber characteristics, the original bale density and the atmospheric conditions prevailing
Carding and prior processes for short-staple fibers
123
at the time the bale is blooming. The specific volume of the bales at the time the bale
plucker is removing clumps affects the size of the fiber clumps and the performance
of the downstream machines. Thus bales should be opened at least a day before they
are put in a laydown.
Of course, the specific volume of the fiber is much higher in transit through the
pipes connecting the machines because the fiber clumps are dispersed in air. It also
decreases during condensation at the collection points at each mechanical feed system.
Every time the material is further opened, the specific volume increases and these
changes have to be taken into account in calculating pipe sizes, feeder speeds, and so
on. The degree of fiber openness also affects what sort of cleaning is effective.
5.5.3 Maintenance of the machine elements
The working elements of the various machines such as beaters, grids, etc., must be
kept in good condition and should be properly set. If the elements in contact with the
fiber become bent, nicked, or otherwise damaged, the fiber is likely to be damaged.
Fibers may collect and clog the machine, and neps may be produced. (Neps are tiny
balls of fiber that degrade the appearance of yarn and show up strongly on the surface
of fabric. Unfortunately neps often dye to a different shade; this emphasizes their
presence and reduces the value of the product.) As was noted earlier, the size of the
machine elements gets smaller as the fiber passes downstream in the process line.
This makes them more vulnerable to damage and increases the likelihood of producing
fiber damage especially if they are not properly maintained and set. Furthermore,
where the relative velocities are high, the abrasion of the metal surfaces increases. A
major purpose of the opening line and card is to reduce clumps of fiber to single
entities. Sufficient working is needed to do this but excess working can only damage
the fiber, produce nep, and remove useful fiber mass. Careful assessment is required
to make sure that there is only just sufficient opening and cleaning.
5.6
Condensation
5.6.1 Feed arrangements of the various machines in the opening line
The fiber stream is carried through some machines purely by the flow of air (e.g. the
axiflow machine in Fig. 5.4(a)) but in others, a mechanical feed is used. The fibers
and the air have to be separated to allow a mass of fibers to be advanced by mechanical
means into the working area of the machine concerned. The first sort needs no
explanation as far as the feed is concerned. The latter does need some explanation.
5.6.2 Accumulation of fibers at the feed
As mentioned earlier, the process of separating the fibers from the air and the
accumulation of fibers on a surface is called condensation.
An example is the weighpan feeder (Fig. 5.5), which uses lattice apron feeds. The
lattice apron is a permeable ‘belt’ on to which fibers are collected and the air passes
through the belt. Another example is a rotating ‘condenser’, which collects fibers on
the inside of a porous drum. Dust from the fibers is pulled through the perforations
to be carried away by suction. A fiber take-off system is applied to the outside surface
to keep the quantity of fiber on the surface at a given level. The openings in the lattice
124
Handbook of yarn production
Fibers out
Beaters rotate
Adjustable
louvres
Fiber in
Grid bars
Trash
(a)
Beaters
Fiber output
Fiber input
Adjustable
grids
Trash
(b)
Fiber + trash input
Condenser
Cleaned fiber output
Dust
+
Air
Gridbars
Trash
(c)
NB Drawings not to the same scale.
Fig. 5.4
Fiber cleaners
apron are fairly large and the size of fiber clump that can be captured is also fairly
large. On the other hand, the openings in the condenser surface are small so that small
clumps and single fibers can be collected on the outer surface.
As mentioned earlier, the process of accumulating many layers of the incoming
fiber causes an appreciable degree of doubling. It will be recalled from Chapter 3 that
the term ‘doubling’ implies a considerable diminution of the unevenness in the material
Carding and prior processes for short-staple fibers
125
Doffer
Level control
Fiber flow
Weighpan
Fiber
feed
Fiber out
Conveyor
belt
Lattice apron
Lattice apron
Fig. 5.5
Return flow
Weighpan feeder
collected on the collection surface. Thus, there is an improvement in the short-term
evenness of the product at that point. (However, after the material passes through the
subsequent machines, the error assumes an ever longer wavelength, and in the yarn
it will be seen as a very long-term error.) The doubling just mentioned tends to hide
the unevenness produced by the opening processes.
Fibers and fiber clumps are attracted to the outer surface of the condenser just
described. Perhaps what is more important, fibers are attracted preferentially to the
thin spots in the fiber mat lying on the screen because more air flows in these zones.
Conversely, there is less airflow to the thick spots and the rate of fiber flow to these
thick spots is reduced. Thus, there is an automatic regulation effect.
5.6.3 Mechanical feeds
The previous paragraph dealt with lattice aprons as condensers, but they also fulfill
another function. They serve as transport systems that move the fiber from place to
place. They often serve as a means to project fiber clumps into a stream of air. (The
airstream is, of course, another means of transport.)
A second mechanical system is to use a pair of rolls to grip the incoming fiber and
feed it forward. The rolls may be pinned or fluted and they have to be set at a distance
apart that will entrap the incoming fiber and not jam up or choke.
A third mechanical feed system is to use fluted or pinned rolls. The rolls engage
a batt of fiber and induce it to slide over a smooth surface. Figures 5.8 and 5.10
relating to chute feeds often used to supply fiber to a card will be shown later.
5.7
The process of cleaning
5.7.1 Philosophy of cleaning
Natural fibers, in the state that they reach the mill, have mineral and vegetable
particles lodged between them. With cotton, there are often seed coat fragments
attached to them. It is difficult to remove some of the extraneous matter without
vigorous mechanical action and without adequate opening. Every time a clump of
126
Handbook of yarn production
fibers is divided, a new surface is exposed from which it is relatively easy to remove
the loose unwanted matter (trash) but trapped or bonded material is a different matter.
The machines in the opening and cleaning line are intended to remove waste but the
amount of waste in the raw material is quite variable. At one extreme (man-made
fiber) the waste is very low. At the other extreme, with some natural fibers, there
might be as much as 10% (or even more) non-lint material. There might also be some
unwanted nep-prone fiber.
As was discussed earlier, cleaning nearly always accompanies an opening function.
As the material is opened, the specific volume of the fiber mass changes considerably.
To maintain an approximately constant mass flow of fiber through that process phase,
the mean velocity has to increase. Imposing such acceleration on the moving fibers
is really saying that the flowing material has been drafted.
Although it seems obvious, the material removed during cleaning must not be
allowed to re-enter the fiber stream. This is because, not only would it be inefficient,
the material removed would contain fibers damaged by the cleaning operation, and
which might cause extra problems in subsequent processing.
5.7.2 Various means of cleaning
There are several ways in which fiber can be separated from trash.
Newly removed fiber is removed from the bale plucker by an airflow running at
perhaps 100 ft/sec (30 m/s). A substantial stream of air carries the fiber to the next
machine. It is usual to have a magnet in the air duct that can remove ferrous materials
from the flow and thus reduce damage to the following machines from these foreign
objects. The magnets are sometimes called ‘humps’ because of their unusual geometric
arrangement. Also some operators use pneumatic separators that throw out nonferrous foreign objects, as sketched in Fig. 5.6. (The sketch is based on sales literature
of Trützschler Gmbh & Co, Germany, to whom acknowledgments are made.) Heavy
particles are thrown out because air is forced to flow in a circular direction; air and
fiber are sucked back into the airstream but heavy particles such as wood and stone
are ejected. It is surprising what is sometimes found in bales of fiber!
In Figures 5.4(a) and (b) it will be seen how the exposed surface of a clump of
fibers is gripped by pins or teeth and dragged over one or more edges formed by grids
to remove trash. The trash drops through the slots and the fibers go on their way. Grid
bars or screens are used in several places in the opening line and in the card. It must
Input = Air + Fiber + Heavy particles
Output = Air + Fiber
Output
Input
Rod
grid
Returned
fibers
Heavy
particles
Fig. 5.6
Heavy particle removal
Carding and prior processes for short-staple fibers
127
be appreciated that with cotton, the trash particles are often attached to the fibers and
it is not easy to remove all of them. As the fiber is worked, trash tends to become
separated from the fiber and it is necessary to have a fairly large number of cleaning
points to be effective.
As was explained earlier, the fibers are usually gripped in these machines by using
saw-like teeth (Fig. 5.4(c)) or pins. The pins can be very large such as those used in
the cleaners shown in Fig. 5.4(a) and (b) or they can be very small as in the case of
pinned rollers used for improved fiber separation. These latter pins are commonly
conical in shape with sharp points and they are more susceptible to damage than the
others. Also the more aggressive action makes fiber breakage more likely and it is
important the fiber clump size be reduced as much as possible before entering the
machines equipped with fine pins. The saw-like teeth are similarly used where the
fiber clump size has been reduced. The saw teeth are used extensively in carding.
Some machines include one or more perforated screens to which suction is applied
for the purpose of removing some of the dust and fine fiber particles. In some other
machines, a rotating condenser collects fibers from an airstream on its outside surface
and dust from the fibers is pulled through the perforations to be carried away by
suction.
Another way in which fiber can be separated from trash is to use the differences
in mass and air drag between a fiber and a trash particle. This is rather like the
method used by the primitive farmer who winnows the chaff from the corn by allowing
the wind to carry away the chaff. The fibers and the trash particles have different
trajectories that permit separation.
A batt of fibers may be beaten and/or vibrated to cause unwanted particles to filter
down through the mass so that the unwanted material can be removed. However, hard,
spiky, or attached particles do not respond to this treatment unless they have been
crushed and/or abraded in a prior operation so that they become detached or less
spiky than they were. The thicker the batt or the more dense the clump of fibers, the
more difficult it is to remove the trash. Effective cleaning in this way cannot be
carried out until the fiber is well opened. Many machines are effective in removing
the dirt on the surface of a tuft; they are less so in removing dirt from the center of
the clumps.
5.7.3 Some examples of cleaners
Pre-cleaning machines are sometimes inserted first into the line to divide the clumps
and removing the worst of the trash. In the case of the ‘axiflow’ machine shown in
Fig. 5.4(a), there is only that overall drafting which is caused by any acceleration of
the fiber stream. However, the tearing apart of the clumps provides drastic local
episodes of drafting. The main emphasis is that of removing trash from the outside of
the fiber clumps. An adjustable louver, or some other air control device, is provided
because it is important that the air pressure should be balanced across the grids. Lack
of proper pressure balance can cause newly released trash to be reintroduced into the
main fiber stream, or more usable fiber is taken out with the trash than is necessary.
If there are local zones of low air pressure, trash can be sucked back into the main
flow. To this end, it is important to make sure that access doors and hatches are
properly closed during operation.
The inclined cleaner shown in Fig. 5.4(b) provides multiple stages of cleaning
with the first stage usually being the most effective. The type of cleaner mentioned
128
Handbook of yarn production
in this paragraph is used at the beginning of the opening line where the size of fiber
clump is relatively large and there is a need to guard against too violent an action
because it would damage the fiber. Each case has to be judged on its merits depending
on the quality level being sought.
The type of cleaner shown in Fig. 5.4(c) is used in later stages where the clump
size has been reduced and the more aggressive parts of the working elements (i.e. the
teeth) are less likely to cause fiber damage because of the openness of the material
being worked. The degree of cleaning has to be balanced against the cost of the fiber
removed, remembering that fiber costs are about half the total cost of the yarn.
5.7.4 Maintenance of cleaning machines
Cleaning machines work in a hostile environment. When working with man-made
fibers, abrasion can be caused by the fiber finish or even by the fibers themselves.
With cotton, much of the dust is silica (or other minerals) from the soil in which the
plants were grown and abrasion of the components of the cleaner is increased by its
presence. Similar remarks apply to other vegetable fibers. In the rare cases where
short-staple animal fibers are used, accumulations of grease or other material tend to
clog the machine.
In the various cleaning and opening machines, distances between co-operating
surfaces have to be set to control the material. As already mentioned, it is desirable
to get as small a tuft size as possible without impairing productivity or damaging the
fiber. As elsewhere, good maintenance is required to make sure that beaters and grids
do not become nicked or worn because such damage can damage the fiber and cause
increases in nep level, both of which reduce the value of the yarn.
5.7.5 Fiber handling
Fiber handling is a term that encompasses not only transport of the fibers, but also the
condensation, cleaning, and control of the fiber flow.
Transport of the fiber is usually by a pneumatic system that carries fiber from the
bale plucker or weighpan feeder to the following machines, between adjacent machines
in the line, and eventually to the chute feed that condenses the fiber before carding.
This involves the use of high volume fans that generate the pressure difference to
create the flow of air. Sometimes the fiber passes through the fan blades, in which
case there is danger of fiber damage and nep creation. Sometimes the fan is at the
receiving end and a condenser is used to collect the fiber and allow the filtered air to
pass to the fan. The design and installation of the air ducts have to be carried out with
care because sharp bends and joints that create turbulence can cause nep. Also jagged
metalwork, protruding screws or the like can create fiber strings that are very difficult
to separate in later processes. A ‘string’ is created by a clump of fiber being caught
up by some projection in the ductwork, twisting in the wind, and trapping more and
more fiber as it lengthens until it looks like a piece of thick string. Fiber strings then
break off intermittently to produce material that is difficult to divide without fiber
damage and is liable to produce chokes.
Machines along the process line often have a sort of chute feed that not only
condenses the fiber but also regulates the flow by mechanical means. A discussion
relating to flow and regulation is given later.
In general, the chute contains some device to measure the height of fiber in the
Carding and prior processes for short-staple fibers
129
chute and some means of controlling the packing density of the fibers in the column.
Vibrating front plates, airflows, and careful geometric design are elements often
employed to maintain the packing density. Once that is maintained, a simple volumetric
control of the fiber feed suffices.
It might be realized that machines with beaters and saw-toothed coverings are
extremely dangerous if left unguarded. In many countries it is a legal requirement to
provide proper interlocked guards that will cause the machine to stop if removed.
Also remember that most of the machines are heavy and the inertias involved do not
permit them to be shut down quickly. Not only are the safety requirements legal in
nature but there must also be a strong commitment by employed and employers alike
in the matter of safety if accidents are to be avoided.
5.8
Intimate blending
5.8.1 The consequence of poor blending
Fibers may be blended in a mill by either mixing them together before carding or by
running several slivers of each sort in a creel of a drawframe or other sliver processing
machine. The former process is referred to as ‘intimate blending’ and the latter as
‘creel blending’. Creel blending might be carried out at the drawframe or the comber
but we must defer that discussion until the next chapter. For the purposes of initial
explanation, it is useful to consider an intimate blend of two fibers, say polyester/
cotton. If the fiber tuft size is too large, the card web will contain streaks of 100%
polyester or 100% cotton, and these streaks appear in the sliver. Even with nominally
similar fibers, streakinees will exist. Certain fiber characteristics will exist as streaks
in the card web just in the same way that the streaks of polyester or cotton had
appeared. Thus it may be seen that clump size in the feed to the card is an important
factor in following stages of production.
Next consider bales of all the same nominal type of fiber. Despite being of the
same type, there are differences in fiber attributes from one bale to another. During
opening, fiber is taken from sets of bales. These sets are set in sequence along the
laydown. Concentrations of fiber from a given bale set might not be completely
dispersed among the rest by the blending. Variations arising from the bale-to-bale
changes in fiber will then appear in the sliver produced.
There has to be a substantial element of mixing in the opening line to make sure
that the fibers from the original bale laydown are properly homogenized. Failure to
do this produces results similar to those just discussed, except that the size and
distribution of the streaks are on a larger scale. Difficulties do not appear until the
yarn is made up into fabric, at which time so-called ‘dye streaks’ and ‘barré’ occur.
Streaks of fiber that do not match the neighboring areas of a fabric can produce an
effect very disturbing to the user. If the streaks are long enough, they appear as bands.
Also if cones of yarn have different properties from others in a lot, this too will
produce bars in the fabric. This is called barré. Such faults in the fabric are common
causes of customer complaint with the responsibility being laid on the yarn maker;
the settlement of such claims can be very costly.
5.8.2 Coefficient of variation (CV) as a measure of blending efficacy
(Efficacy is used in the subheading rather than efficiency because the latter is difficult
to define, as will be realized from the following text.)
130
Handbook of yarn production
Before launching into explanations concerning variation, it might be useful to
define some textile measurement terms. The fiber attributes quoted in Tables 5.1 and
5.2 are commonly used in the cotton industry and they are listed below.
MIC =
Micronaire (a measure of fiber fineness)
This is an old measure of fiber fineness analogous to linear density, which
is still widely used in the cotton industry. The values used to be quoted in
the unlikely units of mg/inch but they are now regarded as just indexes. It
is really a measure of permeability of a wad of fiber in a specified enclosure
as described in ASTM standard D1448.
UHM = Upper half mean length (inches)
The population of fibers in a sample may be divided into those longer than
average and those shorter. The short fibers contribute little to the strength
of a yarn. The long ones make more than a proportionate contribution to
strength. The mean of these long fibers yields a single figure of merit that
gives an idea of what useful length is available in the lot of fiber that was
sampled.
STR = Fiber tenacity (gf/tex)
This is an old standard definition of normalized fiber ‘strength’ or ‘tenacity’,
which is still in use. The new standard is in terms of mN/tex. A tenacity in
gf/tex is multiplied by 9.81 to get it into mN/tex.
ELO = Fiber elongation at break, %
Fiber reflectance determined by a Nickerson-Hunter colorimeter according
Rd =
to ASTM standard D2253
SFC = Short fiber content (%). The length of these fibers < 0.5 in (12.7 mm)
+b
= A measurement of fiber yellowness determined by a Nickerson-Hunter
colorimeter according to ASTM standard D2253
CGRD = color grade, which determines the grayness of the fiber
Area = Percentage of a test surface covered with trash removed from a sample of
cotton under standard conditions.
If a blend were perfectly homogeneous, there would be no variation in the fiber
attributes over any number of samples. Clearly the blend cannot be homogeneous if
any, or all, the various fiber attributes such as micronaire, length, etc., vary over the
set of samples. Variations in fiber attributes tend to be independent; the value of CV
of one attribute is not necessarily reflected in the others. Often, color grade and short
fiber content of the sample have very much higher CVs than the rest of the attributes.
The blend is usually significantly worse concerning short fibers than the upper mean
length of the fibers (UHM). Most notable is the tendency for most CVs to decrease
in the opening line and then increase again in subsequent processing. In the opening
line there are large drafts but there is always a compensating amount of doubling in
the mixer, chute feeds, and the like. There is also removal of some undesirable matter.
In the case cited in Table 5.1, it will be noted that the trash and dust levels fell
markedly in the opening line as, of course, they should. There is often a slight rise in
CVs between the bale and card sliver. In post-carding stages of drawing, the draw and
doubling ratios are usually roughly equal and there is little removal of material even
though the CV might be relatively high.
Nevertheless drawing still tends to reduce the trash and dust content but it often
causes slight increases in the CVs for the other fiber attributes. Such a result was
found by El Mogahzy [2] in an industrial case study. He also found that variations in
Carding and prior processes for short-staple fibers
Table 5.1
131
CVs of fiber attributes at various stages
Bales
Chute feed
Card sliver
First drawing
Second drawing
MIC
UHM
STR
ELO
Rd
Trash/g
14.4
2.4
3.1
2.1
2.9
5.0
1.7
1.3
1.4
2.1
8.3
5.5
3.8
3.3
3.5
12.1
6.3
5.0
5.5
6.2
6.3
1.1
1.2
1.6
1.5
87
16
28
24
24
Dust
73
51
28
27
8
color grade were highly significant in the particular case (see Table 5.2). Space
precludes inclusion of all the data available. It can also be seen that the within- and
between-bale CVs were comparable and that implies that control of the bale laydown
quality is imperfect if no account is taken of the within-bale variance. Normally bale
selection is based on two or three samples per bale and this may be insufficient in
some cases.
5.8.3 Intimate blending
As was mentioned earlier, the first step is to assemble a bale laydown, and the
number of bales in a laydown is usually determined by the operational need to run
without replenishment for a round period (commonly 24 hours). The physical
arrangement is that which is best suited to enable the bale plucker to remove fibers
layer by layer. Once prepared, the bale plucker is set in motion and starts a flow of
fiber into the opening line. The cutting head usually moves to and fro across the top
of the laydown and the height of the head above the ground is progressively decreased
after every traverse. The depth of cut is thus determined. The depth of cut and speed
of traverse of the bale plucker decide the degree to which the material is separated
into tufts. Both of these parameters affect the productivity of the machine and the size
of the fiber clumps generated. There are other schemes for taking fibers from the bale
supply but space precludes further discussion.
Once a stream of fiber clumps is established, there are several ways to homogenize
the flowing material. Apart from the mixing created as the fiber stream passes through
each of the machines already described, there is normally a blending machine, whose
main function is to homogenize the flowing fiber. A typical machine of this sort is a
sandwich blender in which fibers are laid in layers on surface AA as in Fig. 5.7(a).
Table 5.2
A selection of CVs of fiber attributes
Bale no.
2
3
4
6
7
8
9
MIC
UHM
STR
ELO
Rd
+b
CGRD
Area
SFC
2.4
1.3
5.0
6.3
4.2
4.1
38.2
106
19.4
2.3
1.4
4.2
6.6
4.9
6.0
43.0
110
15.4
2.1
1.3
4.2
7.8
4.4
4.9
35.8
99
14.9
2.2
1.5
4.7
5.3
4.4
5.0
38.4
104
14.4
1.8
1.2
5.1
6.0
4.3
5.1
38.6
131
14.6
2.1
1.2
3.9
5.3
4.2
4.3
42.1
132
13.2
2.0
1.2
4.4
7.4
4.6
4.5
46.6
133
12.8
CV (%)
between bales
3.3
0.7 in
3.1 g/tex
4.7%
4.6
4.8
–
–%
11.1%
132
Handbook of yarn production
Output fiber flow
B
A
Input fiber flow
A
(a)
Cell number
1
2
3
4
Fiber input
Cells
Feed rolls
Beaters
Fiber
output
Conveyor belt
Trash
(b)
Fig. 5.7
Blending machine
The fibers flow to B where they are removed perpendicularly to the flow line across
the many layers accumulated. This provides a means of ensuring reasonable homogeneity
in the fiber blend over a certain mass of fiber within the output. If there are variations,
which are very long, they overwhelm the blender and it can only ‘smear’ the boundaries
of the changes. Such a situation certainly can occur when introducing a new type or
merge of fiber. Some blending machines do not work as effectively as the one just
described. In a common type, a series of cells is filled from the top to produce
roughly horizontal strata that are often about 0.5 inch (≈ 12.7 mm) thick. Pairs of feed
rolls empty the cells of stock and deliver the material to a conveyor belt below. The
belt discharges the fiber clumps into an airstream that carries them to the next
machine. If the strata remain roughly horizontal across all the cells and all the feed
rolls were turned off and on at the same time, there would be little blending. In fact,
the cells are not emptied uniformly and each cell should deliver fiber at different
times. In Fig. 5.7(b), cell 2 is being emptied causing the strata of fibers within the cell
to move relative to the others (NB the diagram shows a four-cell machine but larger
numbers of cells are available). This causes blending by taking vertical slices as
described earlier. The positions of the strata marked Z show the effect that is obtained.
Carding and prior processes for short-staple fibers
133
The volume in a cell is limited and the amount of mixing is spread over a limited
volume. Again, very long-term changes in blend may not be smoothed.
The question of recycling waste arises. Klein [3] quotes maximum percentages of
recycled waste fibers as 5% for ring spinning; also between 5% and 20% for rotor
spinning according to count. Hard waste (e.g. roving waste) is often returned to lower
grade products after some disintegration process where the recycled intermediate
product is reduced to more or less separate fibers. Some spinners recycle their waste
through weighpan feeders to give an even flow of waste rather than baling the waste
and reintroducing it into the bale laydown. This is effective in stabilizing the waste
percentage in the fiber stream.
5.9
Fiber flow
5.9.1 Fiber flow control
The crudest of controls is a level switch regulating an earlier but neighboring machine
in the line. When the fiber in the chute rises above the set point, the delivery from the
prior machine is cut off until there is a demand for more fiber. More sophisticated
versions might involve two-speed delivery rates to give a smoother control, but these
are more expensive. Of course, there is always a need to cut off the inflow where
choking is imminent. Continued supply with inadequate removal causes material to
become packed in the ducts and/or machines. Such a disruptive event is a financial
burden to clear because of the human effort needed and the idle time of the machine.
Control of the fiber flow becomes an important matter if the yarn produced is to fall
within proper tolerances of fiber composition and linear density. One form of control
uses an electric eye, mechanical feeler or some other device to measure the height of
fiber within a confined space. When the height exceeds the set value it restricts, stops
or diverts the fiber supply until the offtake has reduced the fiber to a slightly lower
level as shown in Fig. 5.8(a). Machine 2 is controlling Machine 1, because Machine
1 is supplying fiber to it. Failure to stop or slow the flow when the hopper or chute
of Machine 1 is too full would lead to a blockage in its entry port or the ductwork.
Since several machines are used in series, similar controls can act to stop or slow the
feed of fiber from the machines elsewhere in the chain. Machines 1 and 2 are drawn
as cleaning machines but they could have been some other devices in the line. Some
machines incorporate an overflow system. If there is an excess of fiber, the excess is
diverted and recirculated to the feed. As was mentioned earlier, this is a rough form
of evenness control but it also acts to prevent choking. This latter is important because
if the machine has to be stopped to remove the choke in the feed, much of the
production is lost during that time. Usually, there are only a few opening lines and
therefore the economic repercussions of a shutdown can be quite severe.
Modern systems use chute feed systems to supply fiber to a card. The chute
contains some device to measure the height of fiber in the chute and a means of
controlling the packing density of the fibers in the column. Vibrating front plates, the
flow of air and careful geometric design are elements often employed to maintain the
packing density. Once that is maintained, a simple volumetric control of the fiber
feed suffices.
It becomes necessary to insert automatic control to regulate the final yarn count
on a bobbin to bobbin basis. It will be remembered that errors in the sliver produce
long-term errors at spinning. Full bobbins are often randomly mixed with the
134
Handbook of yarn production
Machine 2
Machine 1
Fiber input
Switch
Signal
Sensor
Feed rolls
Control line
Fiber output
Level switch controls feed roll of Machine 1
(a)
Material flow
Fiber input
O
S
F
Fiber output
Signal
Off–slow–fast controller
(b)
Input
sensor
Drafting rolls
Output
sensor
Reference
signal
Comparator
CPU
C = Speed measurement and control
NB Signal flow shown in black and material flow in gray.
(c)
Fig. 5.8
Control systems and chute feed
result that material from bobbins of ‘thin’ yarn might appear next to material from
bobbins of ‘thick’ yarns in the fabric.
Devices to control the level and densities of the fiber tufts are necessary to give
good carding. The fiber level in the feed and fiber packing density of the fiber in both
the reserve box and the main chute must be controlled within strict limits for successful
operation. There are many forms of chute feed but space only permits the inclusion
of one example (Fig. 5.8). This example has been chosen because it shows the need
for careful control to ensure uniformity of mass flow in the supply of a fleece of
uniform linear density and openness to the card.
Carding and prior processes for short-staple fibers
135
5.9.2 Control and autoleveling
Simple controllers are found in the opening line, which are little more than on/off
switches. They stop the feed when the textile material in a receiver in the flow line
(such as a chute) reaches the set point. The set point is the desired level; the user
normally adjusts it for the given conditions. If the material level reaches the off
position in the receiver, it shuts off the supply and when it drops to the on position it
starts up the supply again. It has to be remembered that each control has a lag time;
shut off and restart does not cause immediate cessation or restart of the material flow.
These lags produce error and if several such control systems are daisy chained, the
cumulative lag can become quite large. The lags can lead to instability in the system
unless it is properly designed and set. A slightly more sophisticated version has two
supply speeds (S and F) as well as an emergency stop O (Fig. 5.8(b)). The slow feed
is at a rate slightly below the normal feed rate and the fast feed is slightly above. The
supply rate thus oscillates only within narrow limits around the mean level and the
system can be more accurate and stable than the on/off version.
At the next level of sophistication, it is possible to measure the linear density of a
strand delivered to a machine. An error signal from this measurement can then be
used to offset errors in the supply system.
Devices like these are sometimes fitted in cards and drawframes; they are called
autolevelers. The signal is an electrical voltage, air pressure or some other means of
conveying information. If the measurement is made on the input and the output speed
is changed, the device is a feed-forward device. The drawback to this method is that
it takes no account of the changes it makes or of any changes in conditions. It has to
work by dead reckoning. Thus if the calibration changes, or some other factor intervenes,
an error is created. A typical means of controlling the linear density is to measure the
output but control the input; this is called a feed-back system. A feed-back system
has the consequence that the results of any change are carried by the material to the
measuring device after a delay, but this can lead to instability. An example is where
a transducer measures linear density of a card sliver as it emerges from the card, and
the measurement is used to determine the feed roll speed. The change in draft alters
the linear density of the sliver. If the variation is perfectly periodic, the system can be
tuned to give excellent performance because the variations can be predicted on the
basis of the history of measurement. In practice, random variations in linear density
are present and an important component of the signal from the transducer is
unpredictable. Some idea of how the problems arise can be obtained by considering
a single thick spot in an otherwise flawless portion of sliver passing the transducer.
The one-time increase in signal strength causes an increased draft for the duration of
the passage of the thick spot through the measurement device. If the thick spot passes
before the system can react, a thin spot is created in the following material. This thin
spot later causes a negative signal and creates a thick spot and so on. Thus not only
is there the original error but also echoes of it in following portions of the strand.
Since an actual signal is a mixture of repetitive and random errors, a feed-back
system can correct some components but not others. If the estimate of the time lag or
the amplitude of correction is wrong, the strand will contain not just the random
errors but some of the harmonic ones as well. It will contain the echoes just discussed.
Thus feed-back systems have to be used with care.
A step toward further sophistication is to use both feed-back and feed-forward
devices together in a combined system as shown for a roller drafting system in Fig.
5.8(c). The diagram shows only an input sensor but there are alternatives, which
136
Handbook of yarn production
cannot be further discussed. Roller drafting is sometimes used on sliver leaving a
card before it enters a can and the draft is varied as part of the control system. A
device compares the actual measurement of linear density with the desired one and
passes an amplified value of the difference (i.e. the error signal) to the central processing
unit (CPU). The computer unit uses algorithms designed to reduce the instabilities of
the feed-back portion of the system.
It is normal to change the back roll speed in a drafting system in a card or
drawframe. If the front roll speed is controlled, the speed of the can has to change
also. However, the mechanical inertia of the coiler mechanism and the associated can
full of sliver, resists sudden speed changes that might be called for by the control
system. Such sudden changes produce heavy mechanical loadings on the coiler drive
system. Consequently, the rate of change of speed of the output system with its high
inertia has to be limited. This is done by smoothing the demand for change by
limiting the rate of change of coiler speed. An alternative is to work with a temporary
sliver storage system operating with a low draft roller system in series with the sliver
take-off from the card.
5.9.3 Weighpan feeders
Some mechanical feeders use oscillating combs that remove excess fiber from the
belt or lattice apron and give a measure of volumetric control of the fiber flow rate.
This is a crude leveling device. More sophisticated feeders of this type are fitted with
weighpan controllers that dump specified masses of fiber on to a conveyor belt as
was sketched in Fig. 5.5. A number of such feeders are used to give close control of
blend proportions; even though the technology is old, many are still in use today.
Modern design favors continuous control rather than the intermittent supply inherent
in the weighpan feeders.
5.10
Carding
5.10.1 The function of carding
Carding is where the last major stages of opening and cleaning occur. It is also where
separated fibers are converted into the rope-like sliver form. The functions involved,
like the other machines described, embrace opening (the division of fiber clumps),
cleaning (even though this function is little used when making sliver from man-made
fibers), blending, and condensation. There is, however, another function involved.
This is fiber orientation. The card is the first stage where the fibers start to be
straightened and get some orientation in a common direction. Thus the two following
subsections will deal with fiber separation and the carding action. This latter action
deals with fiber straightening, orientation, and a degree of condensation as well as
the other functions. Following this, doffing (the removal of fiber from the cylinder)
will be considered. Doffing entails a considerable condensation process as well as the
conversion of the fibers from a sheet-like form to the rope-like one called a sliver.
Cleaning will also be given a special subsection. Apart from removing short fiber and
trash, the card also has the task of removing more nep than it creates. Other aspects
of carding will also be considered.
In reading the following it should be appreciated that the speeds are high for such
a large cylinder (therefore care has to be taken to keep the cylinder true and balanced).
Carding and prior processes for short-staple fibers
137
Also, the surface speed is high and the teeth tend to pump a considerable amount of
air and it is important to keep the moving surface covered, where possible, to prevent
disruptive turbulent air currents from forming. Further, the surface has to be kept
covered for safety reasons.
The design of the card developed in the nineteenth century. According to Gunter
[4] there was little basic change over the next century except for the introduction of
‘revolving flats’ that move slowly over the surface of a rotating cylinder. The word
‘revolving’ does not mean that the flats revolve about their own axes but merely that
they move around a specified path. The teeth on the active elements have to be very
fine because they have to be capable of handling single fibers. The order of magnitude
of a typical dimension is 0.1inch (≈ 2.5 mm). It also means that they are vulnerable
to damage from foreign objects. This sets the tone for a discussion of cards. A sketch
of a short-staple card is given in Fig. 5.9.
5.10.2 Fiber separation
The main feed roll advances a batt of cohering fibers and a thick fringe of these fibers
is combed by the teeth of the licker-in as shown in Figures 5.9 and 5.10. This results
Moving flats
Flat cleaning
Fiber input
Trash
Screen
Feed plate
Licker-in Cylinder
Doffer
Crush rolls
Sliver
output
(a)
Fiber to flats
Licker-in
Feed roll
Cylinder
Fiber
feed
A
Feed plate
Trash
Mote knife
(b)
Fig. 5.9
Short-staple card
Trash
138
Handbook of yarn production
(a)
Fiber feed from overhead duct
Feed rolls
Beater
Fiber level control
Fiber to card cylinder
Chute
Feed roll for batt
Main feed roll
Fiber batt
Feed roll
Licker-in
X
Y
Feed plate
Enlargement of drafting zone
(b)
Fig. 5.10
Card feed
in a major separation and acceleration of the fibers, which implies a major overall
drafting stage. Numerous further localized drafting episodes take place between the
cylinder and the flats, which drafts clumps or tufts of fiber and reduces many of them
to individual fibers before the process is complete. (Many dictionaries define a tuft
as being anchored at one end but some textile authorities regard a tuft as a small
aggregation of fibers. In this context, the aggregation of fibers is anchored temporarily
in the card clothing and it might be felt that it is a more appropriate word to use in
carding than ‘clump’.)
As mentioned, the processes of drafting in a card cause some fiber orientation.
Some of this orientation is retained because of the restraints provided by the proximity
of the elements carrying out the process. The orientation is far from perfect but fibers
within the tufts are no longer disposed in random fashion.
Considering this opening function, assume that the clump entering carding machines
is 0.1 lb in size (approximately 107 fibers). Let these clumps each be first divided
into 0.05 lb portions. Next, let the 0.05 lb portions be divided into 0.025 lb tufts and
so on. There would have to be 24 stages of division (224 = 16.8 million) before the tuft
Carding and prior processes for short-staple fibers
139
would be reduced to single fibers. This is a theoretical case and it is much more likely
that there would be large and small portions at each division. The total number of
stages of division would then be dependent on reducing the large portions. Assume
that, on average, each tuft is divided into 4/5 and 1/5. It would now take more than
73 stages of division ((5/4)73 = 11.8 million) to reduce it all to single fibers. The point
of these very approximate calculations is to show the necessity of many stages of
division and redivision. (Note: 0.1 lb ≈ 45 g, 0.05 lb ≈ 22 g, 0.025 lb ≈ 11 g.)
Attention is drawn to the multitude of flats on the top of the machine. As mentioned,
the licker-in removes relatively small tufts of fiber from the ingoing fiber batt, partially
cleans them (if necessary), and delivers them to the main cylinder which carries them
to the first flat. Many of the clumps or tufts are caught by the wire on the flat and a
shearing action causes tufts to be torn apart. (The term ‘wire’ is used to describe the
teeth on cards and some opening devices.) Most of the fibers initially caught by the
flat are returned to the cylinder. Many of the divided tufts and the remaining undivided
tufts are then temporarily caught by a second flat and are torn into smaller tufts. The
process continues in this fashion until the fibers are almost completely separated and
lie as a web on the surface of the main cylinder as it leaves contact with the last flat.
Some 40 flats are needed in the carding zone to obtain the desired fiber separation
and comparison may be made to the earlier calculations. Non-moving carding segments
may be interposed between the licker-in, flats, and doffer to give a greater carding
action. The segments carry fixed wire and are placed close to the moving cylinder
wire; an action occurs there that is similar to that between the flats and the cylinder.
5.10.3 The carding action
Two alternative arrangements of the carding elements exist; in one, moving flats cooperate with the cylinder and in the other, fixed plates or segments are used. In the
first case, some 40 flats are linked together and move slowly over an arc of the
rotating cylinder. The surface speed of the cylinder is usually in the range of 500 yd/
min (≈ 457 m/min). There is a small clearance between the teeth, the setting of which
can be varied from 0.008 to 0.02 inches (≈ 0.2 to 0.5 mm) according to the fiber being
processed. Thus the shear rate is very high and a tuft of fibers caught by one set of
teeth is wrenched apart by the opposing set. There is very little time for the fragments
of the tuft to relax until the next division is applied to them. Consequently the fibers
within the tuftlets retain some orientation in the direction of shear, i.e. in the direction
of movement of the surface of the cylinder. This permits a carding action. Portions of
such systems of flats are shown in Fig. 5.11(a).
It is possible to replace the flats and their cleaning apparatus by a simple curved
plate with fixed teeth (or even with a roughened surface) when carding clean fibers
of relatively even length. Some designs exist in which trash can be evacuated from
between the segments by interposing small wedge shaped plates that deflect the flow
of air (Fig. 5.11(b)). The distances between the tips of the teeth on the fixed tops and
tips of the teeth on the cylinder (i.e. settings) have to be carefully adjusted. Also care
has to be taken with trash evacuation systems to ensure that they do not choke. Such
chokes might not be detected immediately but cause deterioration in quality that
might not be diagnosed in the early stages.
Worn teeth (Figures 5.12 and 5.13) give trouble and it is customary to test the card
output for nep on a regular basis to provide a control. When the nep levels exceed a
level determined by experience, grinding or rewiring become necessary. Excessive
140
Handbook of yarn production
Flats move slowly
Cylinder
(a) Flat-top card
Air + Trash
Air
Cylinder
Tooth sizes exaggerated for clarity
(b) Fixed-top card
Some card flat arrangements
Fig. 5.11
Flat wire
X
α
Sharp teeth
W
Cylinder wire
X
Setting
β
(a)
Worn teeth
A′
A
B′
B
Clearance between teeth increases
from AB to A′B′ as teeth wear
Nep rolls due to action of blunt teeth
(b)
Cylinder
Wire
Section X–X
(c)
Fig. 5.12 Card wire and wear therein
Carding and prior processes for short-staple fibers
(a) Correctly
ground
(b) Worn
141
(c) Over–ground
Fig. 5.13 Microscopic views of card wire
wear as shown in Fig. 5.12(b) would require regrinding to remove the metal between
AA’ and BB’ before an adequately sharp edge could be attained. (Of course, the
clearance would be restored to AB but the tips become wider.) The teeth are case
hardened and consequently there is only a limited number of regrinds that can be
carried out under normal conditions before rewiring becomes necessary. Case hardening
means that the body metal has a thin skin of harder metal. For the extreme case
portrayed, it would then be likely that the case hardening had been ground away, in
which case there would be very rapid wear when the wire was put in to service again.
Also it would be questionable whether a sufficient degree of fiber penetration could
be achieved with the wide tooth tips. In regrinding, too heavy a cut with the in situ
apparatus used to grind the tips of the wire causes burrs to form (see Fig. 5.13(c) for
a view as seen with a pocket microscope). This condition might give good nep
performance at the start but the performance deteriorates rapidly thereafter. If problems
persist, it might be time to investigate other designs of card wire.
The fibers leaving the flats on the surface of the cylinder are sometimes exposed
to another carding and/or cleaning process. Carding segments somewhat similar to,
but larger than, the flats carry out the carding at this stage and further cleaning may
be carried out by installing a knife edge with proper air pressure control and waste
removal facilities. A cleaning edge is an effective way of removing pepper trash but
care has to be taken to monitor the condition of the knife edge. Hard particles and
abrasive material tend to nick and wear the vulnerable edge that then creates nep and
causes operational problems.
Merényi [5] reported the sensitivity of the plate-to-cylinder and flat-to-cylinder
settings. With a 0.008 inch flat setting the mass of flat strip removed increased by
150% as the plate setting was changed from 0.017 to 0.019 inch. The work was
probably carried out with wire1 clothing but it still has some relevance. (Note: 0.008
in ≈ 0.2 mm, 0.017 in ≈ 0.43 mm, 0.019 in ≈ 0.46 mm.)
The reason can be imagined when it is realized that the ingoing nip of two large
cylinders rotating in proximity creates a considerable pressure especially along the
line A–A in Fig. 5.14. Unless the pressure is controlled and contained, it tends to blow
out in the direction of the gray arrows and carry dust and lint with it. There is low
pressure under the cylinder/doffer nip and the flow of air from the high to the low
pressure zone affects the fiber orientation in the fiber transfer zone. The air pressure
gradient in this zone affects air leakage as well as the fiber transfer between the
cylinder and doffer.
1 Modern use of the term ‘wire’ refers to saw-like teeth but originates from the use of wire
embedded in a material fixed on the surface of the cards used in the nineteenth and the first half
of the twentieth centuries. In this particular case, wire refers to the original meaning.
142
Handbook of yarn production
Cylinder
A
A
Air
Air
Doffer
Detaching roll*
Sliver formation
Vc
Vd
Trumpet
VC > Vd
Sliver
Enlargement of fiber transfer zone
*The toothed detaching roll shown is one of several
alternative means of removing fiber from the doffer.
Fig. 5.14
Cylinder doffing
Not all fibers are removed by the doffer and many of them recirculate. Grosberg
and Iype [6] carried out an experiment in which the doffer speed was varied sinusoidally
and it was found that the amount of recirculated fiber varied similarly provided that
the frequency of oscillation was not greater than about 1 per second. This was not a
recipe for a practical operation but a means to determine how much the fiber recirculates
and how much is taken off by the doffer. The work showed that there is a mechanism
of fiber storage on the cylinder that might have importance in blending and doubling
within the system. In other words there is condensation on the cylinder caused by the
numerous layers of fiber being collected there.
5.10.4 Doffing
Modern cards remove the web from the doffer by a so-called wire-covered detaching
roll of small diameter followed by a pair of smooth control rolls. Sometimes crush
rolls are used to crush seed coat fragments so that they can drop out of the fiber
stream. The web is then gathered together and passes through a trumpet or condenser
which converts the web of fibers from the doffer into a rope-like sliver. (Note: this
sort of condenser is quite different from the ones that collect fibers and extract dust.)
In modern machines, the fiber flow is assisted in the transfer by aprons that condense
the web to an intermediate condition before passing through the trumpet and calender
rolls. A sketch of the major parts of a doffer system, but with the belts removed for
clarity, is shown in Fig. 5.14. Should the fiber take-off system be improperly adjusted,
it is possible to produced cored slivers, which are dense or entangled in the central
core. Such cored slivers are difficult to draft in ensuing operations.
Calender rolls press the fibers together to give the sliver added cohesion and then
the sliver sometimes passes to a drafting system, which adjusts the linear density.
(Automatic control of the linear density of the sliver output is common today.) The
Carding and prior processes for short-staple fibers
143
sliver is coiled as it is put into a storage can and the device that does this is called a
‘coiler’. The can is used to transport the sliver to the next operational stage.
Returning to the main cylinder and doffer, the teeth are so angled and the distances
so adjusted that the main cylinder gives up the fiber easily and the doffer collects
some of it. The condensation occurs in the fiber transfer zone (shown enlarged in the
diagram). It might be noted that the faster cylinder speed tends to ‘brush’ the fibers
on the doffer with the consequence that the sliver emerges with a predominance of
trailing hooks.
In older cards, the web between the control rolls and the trumpet is in a free
triangular shape with the fiber being withdrawn directly from the surface of the
cylinder. The open web is helpful in judging the quality of the web. The web is
removed from the doffer by a vibrating comb. Sometimes the web passes between a
pair of crush rolls to break up seed and trash particles and the crush rolls might
replace the detaching rolls. The crush rolls are set with their axes not quite parallel
so that the inevitable deflection in the center of the crush rolls will cause the nip line
to have almost constant pressure along their length. Crush rolls should not be used
with pressure sufficient to damage the fiber and they should be avoided when processing
sticky cotton. Newer cards use different means of removing the fiber from the doffer.
Sometimes belts are used to carry the sliver to one side of the cylinder; sometimes the
sliver is ‘rolled’ and sometimes other means are used, but the principles discussed
here remain the same.
For a reasonable output per card, say 100 lb/hr, and a thin web of fibers on the
main cylinder, it is necessary to have a high surface speed. A cylinder speed might be,
say, 142 r/min or roughly 500 yd/min at the surface. If the sliver were delivered at
65 grains/yd, the sliver delivery speed would be 7.5 yd/min. The difference in speed
illustrates how there is a condensation effect at the doffer. Even though the overall
draft in a card is, say, 100, the draft at the licker-in might be 500 with the speed of
the fiber stream varying accordingly. The data are given merely to give an idea of
scale. (Note: 100 lb/hr ≈ 45 kg/hr, 37 yd/min ≈ 33 m/min, 500 yd/min ≈ 457 m/min,
65 grains/yd ≈ 4.6 g/m or 4.6 ktex.)
The cylinder surface speed is higher than the corresponding speed of the doffer.
This speed difference means that the web is condensed on the surface of the doffer
and the web is usually several times as thick as it was on the main cylinder. Also, it
has to be converted from a thin sheet to the required rope-like configuration. The
normal way to do this is to remove the web from the main cylinder by a doffer whose
surface speed is less than that of the cylinder (see Fig. 5.14).
5.10.5 Fiber cleaning in carding
Since man-made fibers need little cleaning and the major short-staple fiber is cotton,
this section is devoted to the cleaning of cotton in carding. Although the cotton gin
is designed to remove seed hulls, some debris from broken hulls is inevitably caught
in the fibers entering the mill. The seed particles are woody and have seed hairs
attached. Sometimes there are waxy, oily or sticky materials present. In seasons of
insect infestations, cotton can contain sugar and insect excreta, which are sticky.
Such stickiness makes carding difficult and it certainly adversely affects doffing.
Very bad infestations can shut down a mill if there is not a sufficient diversity of
sources of cotton in the mix supplied in the laydown. The only reliable solution to this
problem is to avoid the sticky cotton.
144
Handbook of yarn production
The fibers are not completely separated from the woody material in the cleaning
machines, nor are they in the carding process [7] as can be seen in the left-hand
picture in Fig. 5.15. (For that matter, even drawing does not completely remove the
trash particles as shown in the right-hand picture). Further, fibers still attached to a
seed particle at this point are more liable to nep once the particle is removed. It can
perhaps be realized how difficult it is to clean cotton in the process; it is not just a
matter of shaking out the dirt.
Cleaning functions occur in three major zones in the card when carding cotton or
other natural fibers. The first is in the region of the licker-in, the second is in the
region of contact between the flats and the cylinder, and the last is under the card. We
take these in turn.
Any fibers or trash combed out from the main fiber supply are carried under the
licker-in over grids or other trash separating devices to the fiber transfer zone. This
zone is at the nip of the licker-in and main cylinder (not shown). According to Gunter
[4], the distance X in the enlargement of the drafting zone in Fig. 5.10(b) should be
no less than 0.5 inch. This is because if it were much larger, the trash and tufts of fiber
would become too deeply embedded in the wire. He also says that if Y is too large,
plucking will occur and fiber clumps will be fed to the cylinder. Presumably this is
really a reference to the nip to nip distance, which would be analogous to the ratch
setting in a roller drafting system. Certainly the result is the same. After leaving the
licker-in, the fibers are carried by the cylinder wire to the zone of the flats. Varga [8]
quotes research from the Shirley Institute that established that the best performing
diameters of the main elements of a card were 10 in, 40 in, and 20 in for the lickerin, cylinder, and doffer respectively. (Note: 0.5 in ≈ 13 mm, 10 in ≈ 0.25 m, 40 in
≈ 1 m, 20 in ≈ 0.5 m.)
The flats that have just been removed from the carding zone are brushed clean and
are later returned to service at the front (or back) of the card. Material removed from
the flats is called ‘flat strip’. Forward moving flats are the most common; in this case
the cylinder motion helps drive the flats and the removal of waste is easier. Where
rearward moving flats are used, they meet the fiber in a clean condition at the front
of the card but they accumulate short fiber and dirt as they continue to the back for
cleaning. The dirty flats are not brought into contact with the cleanest fibers. The
wire on the flats retains sufficient amounts of short fiber to clog them (i.e. load)
fairly quickly unless the fibrous material is removed. Thus it is common to brush the
flats continuously to clean them.
Not all the fiber on the cylinder is removed by the doffer; a fairly thick film of
fiber is carried under the card back to the licker-in. It is normal to place screens,
grids, and other cleaning devices under the card between the doffer and the licker-in
to help control the airflow around the cylinder and contribute to the overall cleaning
From card web
From
drawframe
Fig. 5.15 Trash in card web and drawframe sliver
Carding and prior processes for short-staple fibers
145
of the fiber. This cleaning would not occur unless some fibers recirculated on the
cylinder. Care has to be taken to avoid the droppings from being sucked up back on
to the cylinder surface. For that reason the doors to the under-card space must be kept
closed. The trash that drops out here is often intermittently and automatically purged
by a suction system. A poorly designed or defective purging system might agitate the
trash under the card to the extent that every time it activates, it causes trash to be
sucked back on the cylinder surface. This can cause periodic episodes of trashy sliver
to be produced. Consequently it is useful to inspect the underside of the cards during
a purge to see if there is any malfunction.
5.10.6 Card wire
So-called ‘card clothing’ is a continuous length of ‘wire’ containing teeth that is
wound under tension on a plain cylinder and the ends are secured. A cross-section of
the wire is L shaped with the upstanding portion containing the teeth. The base forms
a foundation in contact with the cylinder (Fig. 5.12(c)) and sets the distance apart of
the teeth across the width of the cylinder. The bottom of the L shape of the wire rests
on the periphery of the cylinder. Careful heat treatment is needed during manufacture
of the wire because the tips of the teeth have to be hard enough to withstand wear
(≈ 1000 Vickers). However, such heat treatment makes the metal brittle and the main
body of the tooth must retain its toughness; therefore the body of the tooth is tempered
(≈ 200 Vickers). Skilled use of a flame is required to give the right distribution of
temperature during the hardening process. The hardened tip of the tooth is only about
1 mm thick even when new, so there is only the possibility of a limited number of
regrinds.
The shape and size of the teeth are altered for different fibers and frequently
different pitches of wire are used on the cylinder and flats. The wire for the licker-in
is always of a coarser pitch than either of them. The wire of the flats is easier to access
and clean and therefore finer wire pitches are often used. Too fine a pitch causes the
wire to ‘load’ (i.e. become jammed with fiber). Usually, the spacing of the wire is
described in points per sq inch or sq cm. An aggressive wire is illustrated in Fig.
5.12(a), which would penetrate the masses of fiber and grip them well, but it might
cause fiber breakage with some fibers. Even more aggressive designs use serrated
wire and large values of α or β. Point populations vary from 250/inch2 (≈ 39/cm2) for
coarse, long synthetic fibers to 1000/inch2 (≈ 155/cm2) for long fine cottons. Less
aggressive designs have smaller attack angles α or β and non-serrated wire; these are
gentler but not so effective in opening the fiber masses. The general comments apply
to the wire on all surfaces, but the values differ. There is a whole range of designs
according to the fiber being processed and the degree of initial preparation. To make
informed choices, the mill manager should consult the wire makers.
As stated before, the clearance between co-operating sets of wire (i.e. the setting) is
also important, as is the point population. A lower point population gives better
penetration of the fiber mass and better opening, but such wire is incapable of separating
the fiber as well as wire with higher point populations. When the wire is reground, the
tips of the teeth are removed to get to a new sharp edge. In so doing the width W of the
tooth ends increases as shown in Fig. 5.12(a); after several regrinds, the tip cannot
achieve good fiber penetration and the card has to be rewired. New wire has a tip that
is almost completely pointed. Allowing heavy wear is undesirable; reasonably frequent
regrinding should be scheduled. The actual period depends on the fiber and the
146
Handbook of yarn production
requirements for freedom from nep. Maintenance of the wire condition is of great
importance to the performance of the whole mill; it is not a matter to be taken lightly.
The choice of wire is an art and experience is very important. Mechanical loading
of the teeth is caused by the carding action and this creates a problem with highly
angled wire. The forces generated tend to bend the tips outwards and to reduce the
distances between the tips of the cylinder wire and those of the flats. A reduction in
clearance of this sort increases the loading and intensifies the problem. The tips have
been known to touch and destroy the whole clothing at a cost of $1000s per machine
affected. Wire makers are very careful to avoid these problems but there are constraints
in the design of the wire. Users should always remember that the danger of accidents
of this sort increases as the settings decrease. Close settings are sometimes used to
decrease the incidence of nep. However, the use of close settings should be closely
monitored to check on their stability. The settings between teeth (typical values are
given in Table 5.3) are measured by feeler gages and sometimes the flats are set to be
slightly tilted so that the entering cylinder teeth meet the widest setting. Consequently
care has to be taken to measure the settings in the correct places.
5.10.7 Airflow within the card
An aspect of airflow relates to the involuntary flow found within machines. When
two parallel cylinders rotate in proximity, pressure differences are generated on both
sides of the nip. If the rotational directions are opposite, as with a doffer and cylinder,
an increased pressure is generated at the ingoing nip and a decreased pressure at the
outgoing nip.
Pressure differences of this sort create an airflow particularly when the surface
velocities are high and the surfaces are rough. An example can be cited with a card.
Lauber and Wulfhorst [9] showed that an upward airflow existed which was directed
upwards towards the ‘nip’ between the cylinder and doffer. The magnitude of the flow
was about the same as the surface speed of the doffer. Other tests confirmed this and
established, at least for the cases tested, that the airflow swept droppings from the
doffer towards the nip. The use of proper covers and enclosures permits control of the
flows of air that otherwise would create defects in the product.
The size of the apertures formed by the teeth in the nip and the degree to which
they are loaded with fiber affects the airflow between the high and low pressure
zones. In the case just discussed, the flow is vertically downward through the gaps
between rows of wire. It is suggested that this airflow is a significant factor in the
transfer of fiber from cylinder to doffer. On the entry side of the nip there is a
tendency to expel fibers parallel to the axis instead of letting them continue unhindered
in the tangential direction. Some fibers remain on the cylinder after passing the
doffer. Considerations of conservation of mass flow dictate the mass of fiber to be
transferred but do not control the fiber population on the wire. If the population is
high, it takes significant time for conditions to equilibrate after starting.
Table 5.3
Position
Some typical settings for a cotton card
Licker-in–
cylinder
Flat–cylinder
back
Flat–cylinder Flat–cylinder
middle
front
Cylinder–doffer
0.007
0.010
0.022
0.024
0.008
0.009
0.005
0.010
0.017
0.034
inches
inches
Carding and prior processes for short-staple fibers
147
5.10.8 Nep control
All fine fibers reaching the card have some percentage of nep, and, of course, this has
to be minimized. The settings of the card are very important in this respect. Too wide
a setting between the cylinder and flats will fail to remove as much nep as normal and
will favor the creation of nep. Too close a setting (say 0.008 in) can lead to the
problem of wire damage. Also wear on the wire widens the settings. Problems with
bearings or ‘bends’ can increase the setting needed to avoid the wires touching.
(Bends are guides on which the moving flats slide.) Normal settings are made by
inserting a feeler gage of the right thickness into the gap and feeling the drag between
the metal of the feeler and the wire tips. If the drag is heavy, the gap has to be
widened and if it is too light, the gap has to be closed. Setting is a skilled art and even
professionals vary in the actual settings they produce. Checks are often not performed
until trouble is experienced. One machinery maker offers an electronic device for
testing the setting while the machine is running. Several transducers are fitted on a
rigid strip that replaces one moving flat. Electrical eddy currents give a signal that
indicates the magnitude of the setting distance and this permits a frequent check of
the settings without shutting down the card. Other settings are made with feeler gages
and doubtless technologies will become available for continuously monitoring these
too. As productivities rise and the use of fine fibers increases, the demand for proper
control will rise. New technologies will be required to make it feasible to work in a
region where very close settings indeed are required. The closer the setting that
becomes practical and the more automatic the adjustment, the better will be the nep
performance of the card.
The condition of the bearings and deflections of the machine parts govern the
minimum setting. The fiber diameter and the loading determine the maximum setting.
If the setting is much larger than one fiber diameter, there is a tendency for the
fiber(s) to roll into a nep as shown earlier in Fig. 5.12(b). This is especially true if
there is a nucleus already there or if the teeth are blunt. Nep is a cause of much
concern, especially when spinning fine fibers. One prime site for nep creation is the
zone between the flats and the cylinder. Also, damaged teeth in any of the card
elements can create neps. Several major items have to be controlled; the wire must be
sharp, without damage, and of the correct design for the job. A control chart of nep
count should be maintained and the limits suitable for the market served should be
established. A typical chart of the nep from a card is shown in Fig. 5.16 [10]. It will
be seen how the nep count trend slowly rises (a regression equation is given for this
portion of the curve). Further, it can be seen how the nep count drops after the card
is reground and reclothed. Periodic examination of the condition of the wire with a
pocket microscope is advised; this helps to catch any mechanical damage that has
occurred.
The wire life is taken by some manufacturers of cards to expire after about
800 000 lb of fiber have been processed by the cylinder or doffer. The life of the
licker-in wire is less and is quoted by the manufacturers to be about 200 000 lb. At
100 lb/hr, this is equivalent to running for 333 days for the cylinder and doffer or 83
days for the licker-in. Changes in fiber or mineral dust content alter these figures
considerably; also the linear density and type of fiber are factors.
Continuous monitoring of nep count is a necessity in a modern mill. The card has
to be taken out of service periodically and the time for this to occur is often judged
on the deterioration of the nep count. The regression equation given in Fig. 5.16
applies to the period before reclothing. It implies a cylinder wire life of about 200
148
Handbook of yarn production
Nep count/sq inch
200
y = 0.3x + 111
150
100
Before reclothing
50
After reclothing
= Regrinding
0
0
50
100
Days
150
Fig. 5.16 Card wire life
days’ continuous use (the graph is no more than a single sample of the normal
deterioration of the wire; other cases can differ greatly). Reclothing refers to the
mounting of new wire on the cylinder and other toothed elements. A wise operator
does not rely solely on the machine maker’s estimate but needs to check the performance
of the particular machines directly. (Note: 800 000 lb ≈ 363 000 kg, 200 000 lb
≈ 91 000 kg, 100 lb/hr ≈ 45 kg/hr.)
5.10.9 Air ducting
When a flow of fiber is distributed to a series of parallel branches, the distribution
system requires approximately equal air velocities at all points in the ductwork. After
an offtake from the system the volume of air flowing onward is reduced; the crosssectional area of the duct has to be correspondingly reduced to maintain constant
velocity. Changes in air speed create a danger of the heavier fractions of fiber being
deposited in the early offtakes. Turbulence becomes more probable when an overlarge
duct is used and while this is good for blending it can cause unwanted agglomeration
of fiber clumps. As was mentioned earlier, rough edges in the ductwork cause strings
of fiber to become entangled and, when they break free, they suffer fiber damage as
they pass through subsequent machines. It is also associated with the generation of
some nep in the system. Most systems work at pressures slightly below atmospheric
to prevent the egress of dust and fiber. Leaks in the system waste energy because
extra, unwanted volumes of air have to be pumped. Also a serious leak can reduce the
flow in the upstream ducts and adversely affect the performance of the machines
served by the starved ducts.
5.10.10 Sliver storage
Sliver is a soft, weak, rope-like strand that is easily damaged by stretching, crushing,
or perturbation of the fiber order. Normal short-staple sliver is rarely wound externally
on a package such as used for stronger strands (because of the danger of involuntary
and uncontrolled drafting by stretching). Rather it is coiled into a sliver can in the
fashion shown in Fig. 5.17 and this also facilitates easy withdrawal. The can is a large
(usually) cylindrical vessel into which the sliver is fed for storage. The sliver is either:
Carding and prior processes for short-staple fibers
149
Central hole
Can
Piston
Spring
Fig. 5.17
Plan view showing
coiling pattern
Sliver can
(a) delivered through a rotating coiler head and is laid in a slowly rotating can or (b)
fed through an epicyclic device to generate a sliver pattern without a rotating can.
There is an increasing tendency to use ever larger cans because this reduces
handling costs. If it costs 10¢ to handle a can, then it costs 1¢/lb to handle a can
containing 10 lb whereas it would only cost 0.2¢/lb if a can holding 50 lb were used.
The reason for using large cans is self-evident but it also leads to space problems
especially in the creel of a following machine.
The action of the coiler leaves a cylindrical hole down the center of the sliver in
the can. Too large a hole reduces the storage capacity of the can. Too small a hole
causes the sections of sliver around the periphery of the hole to be tightly packed,
which produces false coiler patterns in the spectrograms of any such sliver tested.
The coils of sliver must be laid with precision to optimize the mass stored and to
prevent damage. To do this, it is necessary for the top layer of the material in the can
to be near the coiler. This is achieved by using a spring-loaded false bottom to the
can. The weight of the sliver depresses the so-called piston and the spring constant is
so arranged that the level of the top layer remains at about the same level. A can with
a bad spring or piston, or an overfilled can, causes the sliver to become crushed and
the crushed portions are more difficult to draft than the rest. The cross-overs of the
coil are regular and tend to be crushed most, so frequently a crushed sliver produces
a periodic error in the material produced. The removal of a coil from a can puts in one
turn of twist. The effect is scarcely noticeable with large cans but it does produce a
slight effect with the very small cans sometimes used for open-end spinning.
5.11
Waste control
5.11.1 Waste generation
Operating machines produce waste products. These waste products may be classified
as reworkable or non-reworkable. The latter are divided into subcategories of (a) nonlint materials removed during processing, (b) fiber unacceptable for the intended
process, and (c) fly removed from the air conditioning or machines. The non-lint
materials include trash, dust, and extraneous objects found in the fiber supply. The
materials in categories (a) and (c) are usually disposed of as discussed later.
150
Handbook of yarn production
Klein [3] quotes the waste percentages for various fibers. The blow room losses
range from 6% for 1 inch cotton, to 3% for 1.5 inch cotton (compared with approximately
15% and 10% respectively of total losses in a carded yarn plant). In buying cotton,
the price includes the non-lint material and has to be considered in valuing the
material. For a plant producing, say, 5000 tons/year, the losses due to unusable waste
in the blow room can be as high as 300 tons/year. The quantity of recycled waste is
higher than this. The value of the waste from a single mill might be measured in
$100 000s/year and the waste that has to be disposed of is measured in $10 000s/year.
Understandably, mills do not clean fibers more than necessary. At one time, it was
thought that the use of an increased number of cleaning machines would improve the
cleanliness of the cotton supplied to the card. It has since been found that, beyond a
certain point, repetition of the same process was ineffective [1] and only induced
unwanted fiber breakage. However, in the working range, the decision about the
amount of cleaning must be the result of a compromise between costs, quality, and
sales. (Note: 1 in ≈ 25 mm, 1.5 in ≈ 38 mm, 1 short ton as commonly used in the USA
≈ 907 kg, 1 metric ton = 1000 kg.)
5.11.2 Waste separation
Cleaning machines are unable to remove the non-lint material without removing
some usable fiber; at times the waste may contain up to 50% usable fiber. Neither are
they able to remove all the non-lint material. Indeed, the removal rates vary from 40%
to 70%, depending on the type of waste, the type of machine, and the running
conditions.
5.11.3 Disposal of non-reworkable waste
Arrangements have to be made to deal with the waste produced. Returns from the
spinning rooms such as pneumafil waste, remnants of sliver and roving, etc., are
usually worked into the flow stream in such a manner to distribute it evenly in the
fiber flow (Fig. 5.18). Pneumafil waste is fiber recovered from the drafting systems
in the ring frames; the fibers are of good length but have been overworked and should
be recycled sparingly. It is rare to exceed about 5% reworkable waste if good quality
spinning is desired and some prefer to keep it down to 3%. Non-reworkable waste is
sometimes sold for uses other than yarn manufacture; more often incineration, controlled
dumping or some other form of authorized disposal is used. Klein [3] points out that
the capital cost of the blow room is less than 10% of the total and a more serious
financial concern is the cost of the waste. All the material that goes to waste has been
paid for at the going price of the fiber concerned. To that must be added a portion of
the running costs of the plant, the costs of baling or otherwise condensing the waste
for transport, the transport itself, and the disposal costs. The waste costs are the sum
of these costs less the resale income, if any. He points out that waste costs can amount
to tens of thousands of dollars a year.
The air discharge from the fans contains dust, fiber debris, and particles that must
be removed before the air can be returned to the atmosphere. Often a two-stage
process is used in which cyclone filters separate the bulk of the waste and cloth filters
carry out fine filtration. The latter are large and are usually installed in a ‘dust house.’
Sometimes parallel systems are used for processing different workroom air discharges.
The size of filtration plant is of the order of 300 tons/year and the associated energy
Carding and prior processes for short-staple fibers
151
Cleaners
Trash
Non-reworkable waste
Slivers to processing rooms
Cards
Bale plucker
Reworkable waste
Line from spinning room
Fig. 5.18
Waste system
costs are significant. Since the air is expensively conditioned, it is normal to return
the clean air to the operating rooms concerned, but an air wash is needed to remove
remaining dust and rehumidify the air. The waste material from the dust house is
compressed into bales or briquettes to facilitate handling. It may then be burnt.
Briquettes are compacted to about 80 lb/cu ft (≈ 1300 kg/m3) and this is about the
same density as a common house-building brick.
Modern day work regulations in many countries apply the rigor of law to ensure
compliance. Thus waste is transferred and collected pneumatically; acceptable designs
of fiber separation, waste baling, and dust house are required. Some fiber may be
reworkable but not useful in the particular mill, in which case the fiber has to be
separated from trash and perhaps de-dusted before resale. The waste is often baled for
disposal; in which case, bale presses are needed. It is helpful to have a bale press for
each type of waste, e.g. comber waste, licker-in droppings, flat waste etc., depending
on the market or use for a given sort of waste.
5.11.4 Blending reworkable waste
Examples of reworkable waste are:
1
2
3
Short fiber (called noil) removed in the process of combing.
Spoiled product from the particular machine.
Waste from which usable fiber can be recovered.
Noil is clean and a saleable item (Section 6.3). Alternatively, it can be blended into
laydowns to supply cards that make sliver for rotor spinning machines.
There are limitations to option (1). Yarns are difficult to recycle, roving less so and
sliver is relatively easy to deal with. Roving has to be stripped from the bobbins
before it can be recovered. Sliver of poor quality and the stripped roving just referred
to may be reworked by making a bale of it, and then placing it in a subsequent bale
laydown. However, care has to be taken to avoid dirty, oily or overprocessed material
and only one or two bales of it should be inserted into any one bale laydown.
Overprocessed fiber behaves poorly during processing and this is why the percentage
of recycled material has to be limited.
152
Handbook of yarn production
Option (3) includes the reworkable waste from post-carding operations but excludes
those mentioned under (2). All the reworkable waste has to be recycled with care.
High percentages also give problems. If waste bales are near to one another in the
bale laydown, a cyclic variation in waste percentage in the fiber stream is produced
that might cause intermittent problems. When the fiber mix is too rich in waste
fibers, production difficulties in subsequent processes are to be expected. By keeping
an even flow of waste, the benefits of recycling can usually be garnered without
undue trouble. Failure to control the waste flow can give most undesirable concentrations
of poor fiber that can gravely affect the performance of the whole mill. Reworked
fibers behave like short fiber because they have been overworked. Fiber crimp levels
have been reduced, lubricants have been removed, and the surface of the fiber has
been damaged, or perhaps broken. Such reworked material in excessive proportions
in the main fiber stream produces drafting errors at every stage and causes increases
in ends down. The result is that running efficiency drops, waste levels rise, and
product quality further deteriorates. Also a portion of the recycled waste drops out
during processing and this is associated with the amount of fly generated (which is
often associated with badly performing mills). However, there is an economic benefit
to recycling a small amount of fiber because the fiber costs are such a large percentage
of the total for the yarn.
5.11.5 Card waste
Waste cannot be ignored in product flow calculations. For the moment, let us assume
that the production efficiency without waste is 100%. If the system produces x%
waste and the throughput without waste is P, the actual output is P(1 – {x/100}). If
y% of the waste fiber is recycled, the regain is {xy/104} and the feed of fiber is
P(1 – {x/100}) + {xy/104}). In other words, the loss in production is {x /102
–xy/104}P. Naturally, the production P falls in proportion to the efficiency and the
actual loss depends on how the plant reacts to the changes in fibers used.
A flat top card produces flat strips that are removed from the flats as they leave the
proximity of the cylinder. These strips are accumulations of short, damaged, and
usually unwanted fibers mixed with some good ones. Settings of the clearances
between the cylinder and the co-operating surfaces are important. As the settings are
reduced below the normal levels, the short fiber content of the fiber carried away by
the flats increases. These wastes are often treated as non-reworkable. Fixed top cards
produce no flat strips but do produce waste. Mote knives, gids, and/or screens under
the cylinder allow waste to drop through, and there is also waste discharged from the
licker-in. This non-reworkable waste probably contains cotton dust harmful to the
worker; therefore it is now the practice to remove the material pneumatically by
automatic means.
Fine trash is referred to as pepper trash. Many cards are now equipped with crush
rolls which calender the emerging card web and crush particles of seed. Some of
these particles are likely to be spiky or attached to fibers before the crushing operation.
After crushing, a great deal of this unwanted material is then able to fall out instead
of being carried forward by the product. It might be noted that there are powerful air
currents generated by the card cylinder, doffer, and licker-in; these air currents can
sometimes recirculate small trash particles and trap them in the material being
delivered.
Carding and prior processes for short-staple fibers
153
5.11.6 Effects of varying the opening and cleaning
If some of the opening machines are bypassed, the card licker-in waste increases and
so does the flat strip. For example, private data showed that bypassing a cleaner in a
production line gave 36% and 4.8% increases respectively; the yarn irregularity was
increased by 1.7% CV and the strength decreased by about 7%. On the other hand,
adding a machine can cause problems too. In another example, an added cleaning
machine caused the licker-in waste to increase by 13.6% and flat waste by 2.5%. This
was because of the fiber damage caused by excessive working. The yarn deteriorated
too, with CV increasing by 1.6% and the strength decreasing by about 5%. Excessive
opening and cleaning can do as much harm as having none at all. A careful balance
is required.
5.12
Safety
5.12.1 General concepts
The machinery used in opening and carding can be very dangerous. Nearly all the
machines use rotating beaters or surfaces with teeth or pins rotating at high speed.
Great care has to be taken by the operators to avoid the dangers of being caught by
the machinery or by the ingoing textile material. Neglect in this area can result in
serious physical injury.
By law, machinery has to be provided with suitable guards to prevent the operator
coming into contact with the dangerous parts of the machines while they are operating.
Interlock switches are nearly always mandated by law and these are designed to stop
the machine if a guard is removed.
The working environment is affected by discharges of particulate matter and noxious
substances. The blow room is particularly vulnerable to discharges of dust into the
atmosphere of the workplace and is, in most countries, subject to regulation.
5.12.2 Safety in the blow room
Historically, the blow room and the carding areas were dungeons of unimaginable
filth. The machinery had many dangerous units with rotating beaters and sharp teeth
moving in unguarded enclosures, leather belts flapped waiting to entrap the careless,
large volumes of dust hung in the air with the result there were many accidents and
many workers became sick. Things have improved very markedly since then but this
is no reason for complacency. Modern machinery has eliminated much of the risk to
the operators (it would be unwise to say that it has been eliminated). However, there
is still much old machinery working in various parts of the world and it is useful to
look at some of the out-of-the-ordinary risks.
A few of the dangers are:
1
2
3
Bales are handled by fork-lift trucks and rules applying to the control of vehicular
traffic in restricted spaces have to be instituted and enforced if accidents are to
be avoided.
The release of the straps from the bales has to be carried out under controlled
conditions to avoid injury from the violent release of the straps when the coverings
are removed.
The bale plucker is a ponderous machine and cannot be stopped in an instant;
consequently control of personnel in the operating range of these robots is important.
154
4
Handbook of yarn production
Many of the machines are very tall and means have to be provided for safe access
to the higher elevations of the machines and ductwork.
As was just mentioned, dust could be a problem; it certainly used to be in cotton
mills and thousands of workers in cotton mills became ill with byssinosis. Machinery
today is required by law to be fitted with means of suppressing dust emissions and the
space in which they operate has to be adequately filtered.
Fortunately modern machinery is enclosed and a variety of safety locks and protective
devices are installed so that the risk to the operators is greatly reduced. Older machinery
requires more scrutiny, adjustment, and strict management to approach the safety
levels required.
5.12.3 Safety in the card room
Most of the warnings relating to the blow room apply here as well. However, there are
more machines that are often closely packed and certain additional warnings are in
order.
The very high inertia of cards poses a particular hazard. It takes some minutes for
the cylinder of a card to stop after the motor is switched off, even if a brake is applied.
The normal interlocks are to little avail if a guard is removed and the operator then
carries out a dangerous act while the cylinder is still rotating. All too many workers
have lost fingers, hands, or even arms by disregarding the rules of safety. The most
dangerous areas are the zones around the feed roll supplying the licker-in and the
doffer. In the first case, never attempt to adjust the batt just entering the feed rolls
because it is all too easy to get caught in the material as it enters the pinch of the feed
roll. In another case, it has been the practice to scoop up the web emerging from the
doffer take-off system and feed it into the sliver take-up device or merely to take a
sample. The danger is in touching the doffer surface with its sharp teeth moving at a
considerable speed.
References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
Szaloki, Z S. Opening, Cleaning and Picking, Inst Text Tech, Charlottesville, VA, USA,
1976.
El Mogahzy, Y E. On-line Monitoring of Fiber Quality: Merits and Challenges, Private
communication, Dec 1994.
Klein, W. A Practical Guide to Opening and Carding, Manual of Textile Technology, 2,
Textile Institute, 1987.
Gunter, J K. Carding Innovations, Josef K Gunter, Durham, USA, 1994.
Merényi, G. The Effect of Reduced Flat Speed on Cotton Carding, Tech University, Budapest,
1957.
Grosberg, P and Iype, C. The Effect of Dynamic Changes in Doffer Speed on Cylinder
Loading. J Text Inst, 82, 4, p 457, 1991.
Criado, J J. TX590 Project Report, N C State Univ, Raleigh, 1977.
Varga, J M J. Technical Innovations in Carding Machines, Tomorrows Yarns, (Ed Hearle, J W
S) UMIST Symposium, June 1984.
Lauber, M and Wulfhorst, B. Non-contact Gauging of the Fiber Flow During Carding and
Drafting Cotton by using LDA, RWTH, Germany, 1995 Beltwide Cotton Conferences, National
Cotton Council, Jan 1995.
Propst A. AFIS In-plant Quality Control, 7th Annual EFS Conference, 1994.
6
Sliver preparation
6.1
Introduction
As previously mentioned, the modern opening line is almost completely automated
and very little labor is required in that department. Between carding, and drawing,
there is less linkage. It is common to use can-changing devices that defer some of the
manual work to more convenient times without impinging on the productivity of the
machine. Such machines are kept in 24 hr/day production between maintenance
cycles except for rare event stoppages, which can only be dealt with by a human
operator. There is increasing use of automatic can movers and bobbin transfer systems
but these are not yet in widespread use worldwide. Linking is always possible, but it
is still early in the machinery development cycle for such devices to obtain universal
adoption.
There are several classes of spun yarn. One class is that of blended yarns that
contain dissimilar fibers, and another is one where the fibers are nominally the same.
An example of the first class is a polyester/cotton yarn. In the second class mentioned,
it is necessary to blend the components because, in fact, fibers from different lots are
not similar (although they are of the same type) but they are never described as
blended yarns. With cotton yarns there are two subclasses. These are so-called carded
and combed sliver. Carded yarns predominate but the more expensive combed yarns
have a good market. Since combing is used for cotton processing and very rarely for
other fibers, the topic has been separated from the other processing and it is discussed
in Section 6.3.
6.2
Drawframe
6.2.1 The concept of drawing
Sliver drawing improves fiber orientation, intimacy of blend, and sliver evenness, as
has already been described. Each drawing head is supplied with a number of slivers
contained in cans. Each sliver comes from a different can, and the combination of
156
Handbook of yarn production
these slivers constitutes the doubling part of the operation. The sliver then is passed
to the drafting zone by a power creel, which transports the fragile slivers without
uncontrolled stretching or other damage (see the left-hand part of Fig. 6.1). The
drafting elements (which perform the drawing operation) then elongate the whole
group of slivers to produce a single output sliver as shown in Fig. 6.1. The sliver
leaving the drafting zone passes through a condenser to a large can where it is stored
in readiness for transport to the next stage of the operation. (This sort of condenser
merely consolidates the passing sliver but produces no doubling effect.)
The drafting system usually consists of three or four pairs of steel rolls and the top
rolls have thick rubber sleeves called cots. (Drafting and drawing are discussed in
more detail in Chapter 3 and Appendix 8.) A linear speed of 500 yd/min is common
and higher speeds, currently up to 800 yd/min, are possible. This has the result that
the productivity of the machines is in the order of 500 lb/hr for each head. For this
reason, very few drawframes are needed and the cost/lb of drawing is low. Fiber
wastage is also low at between 0.5 and 1%. A creel containing from four to eight
feed slivers is used. The linear density of the delivery sliver ranges from 40 to
80 grains/yard. A sliver is often passed through two drawframes; the first passage
is called ‘breaker drawing’ and the second, ‘finisher drawing’. (Note: 500 yd/min =
457 m/min, 800 yd/min = 730 m/min, 40 grain/yd = 2.8 ktex, 80 grain/yd = 5.7 ktex,
500 lb/hr = 227 kg/hr.)
6.2.2 The draw zone
Loading on a drawframe roll is high because of the mass of fiber being processed. For
this reason it is not practicable to use aprons and special care has to be taken to use
the correct bottom roll fluting, hardness of rubber on the top rolls, and settings.
Typical types of fluting are spiral or axial. Another factor is the dust emission, which
tends to increase with speed. Machines are now designed to removed dust and fly.
The accretion of fly and contaminants, as well as the progress of wear on the elements,
has to be monitored carefully. Errors can have a profound effect on later processes.
One faulty component can produce a large amount of substandard material even in a
short interval. The top rolls usually have rubber hardnesses that vary between 70° and
90° Shore, to control wear and performance. Buffing of the cots has to be carried out
with care to avoid roughening the rubber, or else lap-ups will become a problem.
When the rubber layer is reduced by buffing to an unacceptable level, the rolls have
insufficient elasticity to control the fibers and the evenness of the sliver delivered
Sliver
condenser
Power creel
Coiler
Drafting
rolls
Input sliver cans in the creel
Fig. 6.1
Output sliver can
Drawframe layout
Sliver preparation
157
deteriorates. Sometimes the rubber surfaces are coated with lacquer, treated with
acid, or hardened by UV irradiation to increase performance, but some treatments
lack durability. Changes to the rubber also produce error; for example, accidental
irradiation by prolonged exposure of one portion of the rubber to sunlight
causes variations in hardness that produce results similar to a mechanically damaged
roll.
6.2.3 Sliver condensation
Sliver leaving the drafting rolls passes through a condenser containing a sharp contraction
designed to produce lateral fiber migration and enhance sliver cohesion (Fig. 6.2). It
then passes through a trumpet, which further condenses it. One or more devices to
measure linear density are usually mounted here. The trumpet should be changed for
differing sliver weights. A rule of thumb is that the throat diameter (in mm) of the
trumpet should be between 1.6√n and 1.9√n (n is the linear density in ktex) depending
on the weight of the sliver. Take-up rolls discharge the sliver into a sliver passage,
which rotates about XX. Sliver is deposited into a can, which rotates at a different
speed and about a different axis, to make the coiled pattern described earlier. There
are also alternative, planetary systems. The sliver trumpet and the underface of the
coiler head get very hot. A hot spot that is particularly hot and wears more than
elsewhere is the exit of the sliver passage marked Z in the diagram. This must be
smooth and properly shaped when spinning polyester or other man-made fibers that
produce oligomers and other materials that can sublime at high temperature. (Sublimation
is the process of a change in state from a solid to a gas.) Any emission of gas at
these hot spots condenses on cooler surfaces to form a hard, crystalline deposit.
These deposits alter the local coefficient of friction and provide sites for fiber
damage.
Exploded view
From drafting system
X
Condenser
Trumpet
Take-up rolls
Sliver passage
Coiler head
Z
X
To sliver can
Fig. 6.2
The machine, of which
this component is part, is
known as a drawframe.
The system of rolls is
usually referred to as the
drafting system.
Sliver delivery from a drawframe
158
Handbook of yarn production
6.2.4 Autoleveling
One of the purposes of drawing is to improve the evenness of the sliver and the
process of doubling is not enough. It is now common to fit an autoleveler on the first
drawframe and the principle of such an autoleveler has already been mentioned in
Section 5.9.2. The autoleveler measures the linear density of the sliver and compares
it to a standard. A signal, proportional to the deviation, is used to cause a change in
draft ratio sufficient to reduce the error signal and correct the error as far as possible.
At present, little more than linear density is used to control the sliver preparation.
Increasingly, this is measured using online measurements from tongue-and-groove,
capacitive, or optical transducers. Often, these transducers are sited at the input and
output of each of the machines between carding and final drawing. Measurements are
also made in the laboratory on samples of sliver.
6.2.5 Fiber hooks
Carding produces hooked fibers, which cause errors in drafting, reduce the strength
of yarn, increase the end-breakage rate, and lead to a general deterioration in
performance. The hooks are pulled out to some extent in the two passages of drafting,
but sufficient are left to make it worthwhile to present survivors to the ring frame as
trailing hooks. Since there is a reversal in hook direction at every transfer (Fig. 6.3),
and since the card produces a predominance of trailing hooks, it is necessary to have
an even number of transfers. This implies that there should be an even number of
passages of drawing. Within reason, the more passages of drawing, the less the
hairiness of the yarn produced. However, four or more passages overwork the fiber
and the normal custom is to use two passages unless a combing process is used. Each
transfer adds a small cost to the product.
6.2.6 Monitoring
An interesting development is the coupling of a computer to the transducers of all the
drawframes, the signals being used as a means of monitoring the performance of the
frames and personnel. The signal is monitored automatically at whatever interval is
selected within the capacity of the system and exceptions to the normal are reported.
The program can be made to print out spectrograms, which can indicate, at a very
Input sliver with
leading fiber hooks
Filling can
Fig. 6.3
Output sliver with
trailing fiber hooks
Emptying can
Fiber hook reversal
Sliver preparation
159
early stage, any mechanical trouble that develops. This is a valuable feature since
normal quality audit procedures might not find the trouble for many hours, in which
time great volumes of faulty sliver are produced.
The drawframe is a favorite place to site an autoleveler because there are fewer
drawframes than cards; there are arguments for using either of these positions and
many operators use both. Initial expense alone is not always a sufficient reason for
omitting such control systems. The levelers have to be set to correspond with the
correct interpretation of the error signal. For example, a capacitive type transducer
produces a slightly different signal for polyester than for cotton. Also, different fiber
stiffnesses can cause a pneumatic trumpet to give a different result. Consequently,
careful calibration is essential for best performance. Furthermore, variations in the
sliver being processed can cause blend inhomogeneities and it is useful to level the
components at as many stages as is feasible. However, autoleveling is no substitute
for good preparation; the effects of variable preparation may be disguised by autoleveling,
only to appear again at a later stage.
Drawing is a very important process stage, and it is used in all forms of staple yarn
production. As has already been mentioned, it serves not only the functions of fiber
alignment, blending, and long-term error reduction; it also serves to smooth out
inevitable differences in card sliver, especially if the routing of sliver cans is carefully
controlled. Drawing is a sort of central clearing house because there are so few
frames needed and a sizable proportion of the total production passes through each
drawframe.
6.3
Combing
6.3.1 An outline of the combing operation
For high quality yarn, an extra process is introduced called combing. The purposes of
combing are to (a) remove short fiber, and (b) improve fiber orientation. Combed
sliver has a ‘silkier’ appearance than card sliver because of the enhanced fiber alignment.
The first stage in this series of processes is lap winding which follows a drawing
stage (Section 6.2.1). One or more passages of drawing are used before combing to
straighten and orient the fiber hooks for best combing performance (Fig. 6.4(a)). A
batch of cans is moved from the transient storage area following drawing to the
creeling area for a ‘lap winder’. Comber lap is then prepared from drawn slivers. The
cans are assembled in a lap winder creel, which transports the slivers to the running
lap winder (Fig. 6.4(b)). Many slivers (often 12) are combined to make a closely
spaced sheet of slivers, which are wound as a continuous layer on to a cylindrical
center (Fig. 6.4(c)). The resulting ‘lap’ might be as large as 20 inches (≈ 0.5 m) in
diameter by 12 inches (≈ 0.3 m) wide and weigh about 45 lb (≈ 20 kg). The lap is
transported to the combing machines where combed slivers are produced (Fig. 6.4(d)).
The sheets of slivers are combed and thereby drafted down to fiber webs. These webs
pass over curved guide plates to the table of the comber where the fiber webs are
layered to form a sandwich. They then pass to a drafting system that restores the
linear density of the strand and converts it to a combed sliver. Combed sliver is coiled
in a can and passes to a transient storage area. The last stage in this series of processes
is the ‘finisher’ drawing using conventional drawframes as shown in Fig. 6.4(e). For
space reasons, the combing machine is shown as having four combing heads but
actual machines have more. The process involves a great deal of doubling, thus one
160
Handbook of yarn production
Carded sliver
(a)
Drawframe creel
Drawframe
Can movement
(b)
Slivers
(c)
Lap winder creel
Center tube
Lap winder
Lap movement
Webs of fiber pass over
curved guide plates
(d)
Webs from each combing
head are layered on top
of the ones from the
preceding head(s)
Drafting
Comber
Can of
combed
sliver
Can movement
Drawframe creel
Drawframe
(e)
Can of
combed
sliver
The drawings are symbolic and the components are not
necessarily in scale nor are all components included.
Fig. 6.4
Drawframe to lapper sequence
Sliver preparation
161
expects the slivers to be more even than usual; further the combing process introduces
a desirable orientation to the fibers that provides the yarns with a silky desirable
appearance. Extraction of short fiber coupled with the good fiber orientation increases
the strength of the yarn made from combed sliver. In normal practice, only fine
cotton yarns are made using the combing process; it will be realized that the yarns are
expensive.
6.3.2 Lap quality
The output speed of a lap winder is up to 120 yd/min (≈ 110 m/min) and it has a draft
of only 2 to 4; consequently there is a considerable doubling effect and very little
drafting error. Within the limits imposed by the capacity of the cans in the creel, the
short-term evenness should be improved.
Theoretically, the CV should be 1/√m of the average value in the input slivers,
where m is the number of slivers in the lap ribbon. However, the variance between the
slivers has to be taken into account. Variation between the slivers can cause some to
be gripped weakly by the drafting rolls and extra errors are created in drafting; it is
important to use even and similar slivers. It is also important that the ribbon of fiber
is wrapped on the lap at the correct tension. Too low a tension produces a soft lap that
is prone to damage in transport and handling. It also increases storage space needed.
Too high a tension makes it difficult to unwind the lap at combing, especially the last
few layers. The ribbon does not part cleanly; there are hairy connections between the
departing ribbon and the remaining cylindrical part. These are known as ‘split laps’.
Klein [1] reports a web doubling process in which the pre-combing drawframe stage
is replaced by a ribbon lap machine that follows a sliver lapper. Web doubling has
many features to commend it and, in the future, it may well appear in other processes,
perhaps even in carding. Many laps are mounted on a combing machine to yield the
same number of comber webs, which are combined by layering and the layered
ribbon is drafted and then condensed to form combed sliver.
6.3.3 The combing process
Figure 6.5 shows the main elements of the combing process. Such machines are
complex and the sketches are meant to extract only the essence of the process. A
common feature is the combing roll, which is shown in different positions in diagrams
(a) and (b). Diagrams (a) and (b) show the left and right parts of the machine. In some
machines this is called a half roll. The combing roll contains one or more segments,
known as combing segments, which have toothed wire or needles to penetrate the
fringes and remove short fibers. Diagram (a) shows a ribbon of slivers that has been
delivered from the lap by the feed rolls A. The ribbon is then nipped by the elements
X and Y, with a fringe of fibers protruding to the right. The diagram shows the fringe
being combed by the combing segment at B. When the combing segment has passed,
and the primary combing portion of the cycle is complete, the nippers are moved
towards the detaching rolls and are then opened. Meanwhile, the detaching rolls F
have reversed and carried back a portion of the fiber web processed during the last
cycles. The newly combed web carried forward by the nippers is now laid on the
returned web to make a piecing D. At this point, the comb E in diagram (b) penetrates
the two layers as the detaching rolls resume their forward motion. The movement
through the comb provides a secondary combing that removes some fibers that escaped
162
Handbook of yarn production
Feed rolls
A
Lap
Nipper X
B
Combing
segment
Combing
position
Combing
roll
Cushion plate Y
(a)
Comb
E
Brush
F
C
G
D
H
Combing
Piecing zone
roll
(b)
Detaching rolls
Combing segment
Fig. 6.5 Stages in combing
earlier removal. The comb is then withdrawn and the nippers are returned to their
original position to start the next cycle. As the combing segment passes underneath,
it meets a brush that removes the noil, which is then removed by a suction system.
Noil is a by-product of the comber; it comprises good and reasonably clean fiber but
it is often used as a component of the fiber supply for rotor spun yarns and some
other products.
The rectilinear motions have to be balanced because modern combers work at up
to 300 nips/min. There are multiple combing positions in a single machine and the
output from each position is combined with that of the others. The doubled web is
then condensed through a trumpet and drawn to form a sliver. Adjusting the draw
ratio permits adjustment of the output linear density. The doubling and drawing
continue at constant speed. Since the detaching roll has an oscillating motion
superimposed on the steady speed, there is a need for a loop G in the individual webs
to accommodate the differences in velocity. In some machines the web is condensed
into sliver before doubling, and in others the webs are laid together sandwich fashion.
If the teeth of the combing elements become bent or damaged, nep and other defects
can be produced. The outermost teeth seem most prone to damage and inspection of
the selvages of the webs sometimes provides an early indication of a need for
maintenance work.
Settings can be adjusted to remove the desired amount of noil. Removal of too
much is an expensive proposition because of the cost of fiber, but removal of too little
reduces the quality of the product. Economics usually decide the issue; the sales
appeal of a combed yarn often resides in the label ‘combed’ and the decision is not
wholly a technical one. Fractionation of the fiber is imperfect and some long fibers
are removed with the shorter ones; also some longer ones are broken. Thus, the
comber is an instrument that usually improves the short fiber content but it does not
bring it to anywhere near zero.
Sliver preparation
163
6.3.4 Comber noil
Parallelization of the fibers in the input, and the linear density thereof, are important
in combing. With insufficient drawing or too heavy a lap, the comber tends to pluck
clumps of fibers from the stock rather than handle single fibers. Uneven slivers tend
to give uneven gripping of the web which allows more plucking to occur than would
be found with good slivers. To avoid overloading the combs, the sliver weight should
be limited, otherwise quality suffers. Many people regard the percentage of noil
removed as a litmus test of quality, whereas it is often the case that more can be
gained by paying attention to the quality of the lap than by increasing the amount of
noil removed. Short fiber is removed to improve the characteristics of the yarn. The
average fiber length in the noil can be varied. Within limits, the more short fibers that
are taken out, the stronger the yarn, and the less hairy and more expensive it is [2].
Figure 6.6(a) shows various regions of a typical fiber diagram before combing. The
section colored black represents the proportion that is almost completely removed as
comber noil and the detachment setting controls the proportion. The section shown as
light gray is almost completely retained and the section shown as dark gray is only
partly retained, the remainder going to comber noil. Several factors influence this
mid-zone including fiber type, CV, and orientation, as well as machine condition and
speed. Normal levels of noil removal vary between 6 and 14%. A fiber diagram of the
noil removed is shown in Fig. 6.6(b) and it will be seen that removal of some longer
fibers occurs; this is unavoidable.
6.3.5 Web layering
There is a doubling arising from the layering of the webs from the individual combing
heads on the comber table. If there are x heads, the theoretical CV is 1/√x of the mean
input value but to get the linear density of the output sliver back to a value compatible
with the following drawframe, there has to be a drafting stage. As will be realized, the
drafting will introduce some error and the improvement in CV is less than might be
expected. The extent of the change in evenness in the combing process depends on
the fiber and the setting of the machine.
Partially retained
Detachment setting
Fiber
length
Feed/Cycle
Retained
Nip of detachment roll
Removed
Number of fibers
Bite of nippers
Fiber
length
(a) Feedstock
Number of fibers
(b) Noil
Fig. 6.6
Fiber diagrams relating to combing
164
Handbook of yarn production
6.4
Creel blending
6.4.1 Basics of drawframe blending
Drawframe blending is used to improve the uniformity of the blend and this applies
no matter whether the slivers being ingested are similar or are completely different.
Slivers of each single component are blended at the drawframe; this avoids difficulties
in carding fibers of widely different characteristics. The idea is to combine a number
of slivers as they enter the drawframe and this very process of doubling blends the
fibers. In the first passage of drawing, each ribbon could be of a specific type of
fiber. In this case, we use a deviation in blend rather than a deviation in linear density
for our calculations. Suppose a spot in one of the slivers has only 40% of fiber A
instead of 50% and the difference in mass is made up by fiber B. This is 80% of the
expected value for fiber A, which makes the blend ratio 40/60 in the bad spot and 50/
50 elsewhere. Assume that the linear density of each input sliver is 4 g/m and there
is normally 2 g/m of both fibers A and B. At the bad spot the linear densities of fibers
A and B in the one bad sliver are 1.6 and 2.4 g/m. The theoretical blended eight-sliver
totals are now (7 × 2) + 1.6 and (7 × 2) + 2.4 g/m giving the proportions 15.6/16.4.
This represents 97.5% of the perfect value for fiber A instead of 80% in the bad spot.
Thus, not only does the doubling reduce the error in linear density, it also improves
the blend evenness. However, as mentioned elsewhere, drawing introduces error that
offsets these gains. It will be noted that in Table 5.1, drawing did not always decrease
the CV and in the second passage the CV usually went up a little.
6.4.2 Fractionation
A card and opening line can separate blend components, especially if the fibers differ
in attributes. For example, the flats in a flat card tend to remove the coarse and short
fibers and the delivery might be depleted of these components irregularly. The
fractionation of the fiber changes the population of fibers and the consequential
variations in populations of fibers affect the blend from place to place in the material
flowing through the system.
The blend for one particular fiber attribute differs from that for another attribute
within the same population. This is because of the many permutations of fiber properties
in the stream of material passing through. A typical industrial performance is shown
in Fig. 6.7 and it will be noted that, not only does the linear density (or ‘sliver
weight’) have a CV but so do all the fiber properties. The CV of the component level
is a measure of the perfection of the blend. Obviously, a 0% value would mean that
the blend is perfectly even. The ‘Uster 25%’ refers to the Uster statistics, which show
the range of worldwide mill performances in respect to evenness of linear density of
the product. A 25% rating means that 25% of all spinners are better than the subject
one. The other symbols are explained in Table 5.1 but special mention is made of the
CV of the short fiber content (SFC) which is usually much higher than the other
values discussed here.
6.4.3 Longitudinal fiber migration
The process of drafting causes fibers of differing characteristics to move relative to
one another during processing. This migration of fibers mixes them and is an unintended
form of blending and it applies whatever input fiber components are involved.
Sliver preparation
165
12
Sliver CV%
Carded sliver
Drawn sliver
8
4
0
Uster Sliver
25% weight
MIC UHM STR ELO SFC
Fig. 6.7 Variability in sliver
Assume that the input material to a machine that drafts the material enters as a
series of sections labeled A, B, C, and D, as shown in Fig. 6.8. The number of fibers
in each cross-section is assumed to be the same. After passing through the machine,
drafting causes longitudinal migration of the fibers and a cross-section of the output
contains fibers with mixed labels as shown. The labels refer to strata parallel to the
material flow. The fiber order in the input is not preserved and there is a mixing of
the input segments. Not all components migrate at the same rate. Mixing will be
biased according to some fiber attribute or attributes. The exact mechanisms are still
unclear at the time of writing. Each output cross-section contains a sampling of
several length-segments entering the machine.
6.5
An industrial case study
A case of a particularly bad laydown in a mill not equipped with a mixer is now
discussed to emphasize the importance of adequate blending [3]. This is not a normal
circumstance and should be considered as a worst case scenario.
The bale laydown consisted of 40 bales and the bale plucker took about 5 minutes
for a round trip. The length of card sliver delivered in the average time for the bale
A B
A B
B
D B
D D
A B
A
C D
C
B
Material flow
Machine
Output
B
B
Input
C
A
C
A C
B
A
D
Cross-sections
Fig. 6.8
Effect of processing on a blend
C
D
166
Handbook of yarn production
CV of fiber fineness
plucker to make one round trip was about 700 yd (≈ 640 m). Tests were made at
10 yd (≈ 9 m) intervals along the sliver and at corresponding times from the bales.
FFTs of the data series were made to show the periodic components of the various
fiber attributes. (FFTs are Fast Fourier Transforms, which convert data using the time
along the X axis to a frequency or wavelength basis. A typical use of the wavelength
basis is the radio spectrum.) To simulate the blending occurring in the opening line,
the data were smoothed over a moving interval of 20 bales. (A ‘moving average’ is a
succession of the averages over successive batches of data and the width of each
batch is termed an ‘interval’. If all the intervals are made the same we refer to a
‘moving interval’.) The CVs of fiber fineness (micronaire) from these tests are shown
in Fig. 6.9(a). The result was primarily affected by the bale-to-bale difference in the
laydown. The laydown contained many bales, the round trip time of the bale plucker
cycle was long, and the mixing volume of the blending system was small. There was
an approximately 700 yd (≈ 640 m) error wavelength in micronaire due almost
entirely to this cause. Tests at other establishments with adequate blending only
showed weak tendencies towards this sort of error. The importance of the particular
variability shown arises because of differences in dye uptake with cottons of varying
fiber fineness.
There are factors other than fiber fineness affecting the dye shade but they showed
little or no effect on the result in this case. It is interesting to look at yarn made
directly from card sliver because this avoids the distortions from yet another production
stage. It was expected that different dye shades would show similar patterns in the
fabric.1 One yard (≈ 0.9 m) of each 10 yd (≈ 9 m) sample of sliver was converted to
Fiber source
Bales
Sliver
Fiber
Color variance
(a)
Color wavelength
400 nm
600 nm
Fabric
100
700
1000
(b)
Error wavelength in equivalent card sliver, log scale (yards)
NB All samples were taken from a single laydown and
no blender was used.
Fig. 6.9
Fiber fineness and dyeability
1 Color is also measured as wavelength, but the unit is nm, that is 10–9 meters, rather than the
thousands of yarns used as the abscissas in the graphs here.
Sliver preparation
167
yarn on a sliver-to-yarn ring spinning machine, and a portion of each sample of yarn
was knitted into fabric. Each product could be directly related to the corresponding
length of card sliver. Figure 6.9(b) shows that the fabric had variations in dye uptake
consistent with movements in the bale plucking head across the bale laydown.
Extremely long-wavelength errors such as those generated in very long bale laydowns,
or large variations throughout the height of the laydown, cannot be completely
extinguished by normal blending devices in the opening line. However, mixers of
adequate capacity can usually smooth the results to an acceptable level.
References
1.
2.
3.
Klein, W. A Practical Guide to Opening and Carding, Manual of Textile Technology, 2, Textile
Institute, 1987.
Pillay, K P R. Text Res J, 34, 663, 1964.
Lord, P R and Rust, J P. Blending as a Systemic Problem, Proc. Beltwide Cotton Conferences.
Nat Cotton Council, Vol. 3, p 1631, Jan 1994.
7
Short-staple spinning
7.1 Ring spinning
7.1.1 Ring spinning and associated processes
A ring-spinning machine is an uncomplicated, flexible, low cost device that is well
established with a wide range of applications. In the past, differences in fiber length
between cotton and wool determined whether the system was regarded as being in the
short- or long-staple category; that view persists even though there are now many
more types of fiber and machines. However, in this book we shall persist with the
original demarcation and, for the present purpose, we may define ‘short staple’ as
covering the range of fiber lengths up to, say, 2 inches (≈ 50 mm). Short-staple
spinning machines may process a variety of fibers, the most important of which are
cotton, polyester, and blends thereof.
Although it was initially developed in the nineteenth century, ring spinning still is
attractive for a wide range of services and is likely to endure for many more years.
The systems described in this section include roving production, ring spinning, and
winding. Roving is an intermediate product made from sliver and it is normally used
as a precursor for yarn. A problem that requires attention is end-breakage in spinning,
roughly half of which arise from faulty roving preparation. This is mentioned to
underline the need to consider the whole production line; concentrating on individual
machines is not sufficient. Since sliver production has already been discussed, we
now continue with roving production.
Essential parts of a roving frame are:
1
2
3
4
A creel, which contains cans from which sliver is drawn to feed the drafting
systems (see Section 6.2.1),
Drafting systems to reduce the linear density of each sliver to that of rovings (see
Section 3.7),
Flyers to twist the emerging rovings.
Individual winders to take up the twisted rovings onto bobbins.
Items (3) and (4) are usually combined as will be seen in the next section.
Short-staple spinning
169
7.1.2 Flyer twisting and winding of roving
The flyer (Fig. 7.1(a)) is ideal for twisting low strength strands such as roving, but it
has a limitation in speed because of the sheer size needed and the associated mechanical
stresses. However, the productivity is reasonable for roving because both the count
and twist are low.
One revolution of the flyer inserts one turn of real twist in the roving emerging
from the drafting system. One turn of the bottom feed roll delivers D inches of yarn,
where D is the diameter of the roll in inches. If the flyer rotates kπD times during a
single rotation of the front drafting roll, the twist is k tpi. The factor k is controlled
by a twist gear, which is part of the gear train connecting the bottom front roll and the
flyer. Since the feed roll speed is fixed, it is also necessary to fix the flyer speed to
maintain an unchanging twist level.
The roving is supported inside the flyer arm or in a slot, and this reduces the forces
acting on it due to centrifugal force. Winding tension is, in part, controlled by wrapping
the roving around the presser arm two or three times. The more wraps, the higher the
tension and the more dense the roving package. In addition to real twist there is also
false twist involved. The rubber grommet provides a surface on which the whirling
yarn rolls to produce false twist between the grommet and the feed roll. This false
twist strengthens the weak twist triangle and reduces end-breakages.
It is not enough to merely twist the strand; it has to be wound on the bobbin and
this requires that the bobbin speed be different from the flyer speed. In some machines,
the bobbin speed is greater than the flyer (one can tell by the direction of the foot) and
sometimes vice versa. The one is referred to as a ‘bobbin-lead’ machine and the other
as a ‘flyer-lead’ machine. In both cases, a so-called ‘warp wind’ is used, and this
means that cylindrical layers of roving are laid onto the bobbins (Fig. 7.1(b)). This
has certain consequences. After the first layer of roving has been laid on the bare
surface of the cylindrical bobbin, it is necessary not only to change the direction of
lay but also to adjust the bobbin speed, because the diameter has just become larger
by two roving thicknesses. After each complete layer is wound onto the bobbin (i.e.
after each lay), the direction of lay is changed and so is the bobbin speed.
This is usually accomplished by means of a pair of opposed cone pulleys and
differential gearing. The cone pulleys are set parallel but in opposite directions so
that a belt connecting them may be moved parallel to the cone axes without being
stretched or going slack. Movement of the belt produces different speed ratios. The
differential gearing is merely a mechanical means of combining two separate input
speeds to give an output speed proportional to the difference of them. In a conventional
roving frame, one input is calculated for the case when the flyer and bobbin speeds
are equal; the other is arranged to give the appropriate wind-on speed. The ‘wind-on’
speed is changed at the end of each traverse by altering the position of the belt on the
pair of cone pulleys just described. In some modern machines, the speed change is
controlled electronically. If the speeds are improperly adjusted, the winding tension
is changed after each layer of roving is placed on the bobbin. If the roving is larger
in diameter than allowed for in the calculation, the bobbin becomes very compact and
difficult to unwind at the next stage of processing. If the strand is wound too slackly,
the package becomes unstable and prone to damage. (If the changes are too great,
there is danger of causing a machine stop due to a broken end.) Thus, selection of the
correct lay gear is important. The precise choice depends on the fiber being used, as
well as the count and twist of the roving. A production unit usually has data, based on
experience, for the best value to use. The drive train controlling this speed change
170
Handbook of yarn production
Grommet to provide
false twist
Feed
α
Threading slot
Bobbin
rotation
(ωb)
Db
Flyer
rotation
(ωb)
t
L
Traverse
Warp wind
Several wraps of
roving around the
presser arm to
control tension
r
D
Each layer contains m coils/inch
(b)
(a)
α
A
B
Diagrammatic representations
Not to scale
(c)
Fig. 7.1
Roving spindle and bobbin
Short-staple spinning
171
contains the lay gear. If an end breaks and the machine is not stopped, these speed
adjustments get out of kilter; further operation is not possible until the machine is
reset.
To give good stability to the roving package, it is normal to progressively change
the length of traverse of the flyer foot with respect to the bobbin. Each successive
layer involves a progressively shorter traverse with the result that the completely
filled bobbin has conical shoulders of about 80° to 100° inclusive angle (which
reduces damage from handling). The setting of the traverse mechanism determines
the slope of the shoulders of the package.
The strand is fed to the twister at a steady pace and therefore the flyer speed has
to be maintained to keep the twist level constant. However, the winding-on speed has
to be varied to match the feed of the roving. The ratio of bobbin and flyer speed1 is:
Ub /Uf = 1 ± 1/τc
[7.1]
The sign in the equation changes according to whether it is a flyer-lead or a bobbinlead design.
The twist is adjusted by changing the twist gear (part of the gear train connecting
the front and back rolls). There was discussion of twist and draft in Chapter 3.
7.1.3 The roving machine
Good descriptions of roving machines are given by Klein [1]. The machines comprise
between 60 and 120 spindles, each containing a drafting system and a flyer twister.
Rotation of the flyers twists the strands and, since the strand is supported within one
of the flyer arms, centrifugal force does not cause it to be tensioned due to ballooning.
It is not possible to rotate such a flyer at very high speeds because of mechanical
design difficulties; speeds up to 1600 r/min are obtained in practice. Bearing in mind
that the productivity of a spindle is a function of the linear density of the strand, it
will be realized that a flyer frame finds its best use in processing relatively thick,
weak rovings that vary between 0.2 and 1.2 ktex.
Input to a roving frame is ‘drawn’ sliver, taken from a can filled in the last drawing
stage. Sliver is normally drafted by the roving frame to roughly 1 hank/lb and it is
then twisted just sufficiently to permit handling before being wound onto a large
bobbin. A roller drafting system is used as discussed in Chapter 3 and in Appendix
8. The top rolls are similar to those used in ring spinning (Fig. 7.2(b)). The spindles
and flyers each share common drives and the bottom rolls of the drafting system are
formed from long steel bars, which are articulated along the length of the machine.
These arrangements enable the gearing to be concentrated in a headstock at one end
of the machine; they also enable the gearing to be enclosed for safety and cleanliness.
The top rolls are similar to those used in ring spinning. Note: 0.5 hank/lb ≈ 1.2 ktex,
1 hank/lb ≈ 0.6 ktex, 3 hank/lb ≈ 0.2 ktex.
The building motion is controlled by the steady upward and downward movements
of the rail containing the bobbins and spindles. The bobbins and spindles are coaxial;
the flyers all operate at the same speed and direction (by custom, Z twist). The
bobbins rotate at a common speed. The best time to set the tension is when the build
1 Relative winding speed = material supply speed, thus (Uf ± Ub)πD = Uf πD/τc which leads to
Equation (7.1). τc is the twist/unit length, Ub is the bobbin speed, and Uf is the flyer speed. It
might be noted that Uf and τc are virtually fixed by the machine design and product, respectively.
172
Handbook of yarn production
Break draft
Rubber-coated
top rolls
Break draft
Main draft
Main draft
X
X
X
Fluted metal
bottom rolls
X
Twist
triangle
Yarn
Short apron
Twist triangle
Long apron
Yarn or roving
(b)
(a)
Wear point
Apron
Front top
roll
X
X
Bottom roll
Yarn
Nose piece
or bridge
Setting
Apron
(c)
The aprons shown are thin reinforced rubber bands, about 1 inch wide,
which move at about the mean speed of the forming yarn. Where they
protrude into the nip of the front rolls, they slide over nose pieces X.
Fig. 7.2
Apron drafting
of the bobbin just begins because, if an end breaks and the machines are not stopped
quickly, these speed adjustments get out of balance and further operation is not
possible until the whole machine is reset. The twister has already been described.
Because the roving packages are large they are usually arranged in two rows (Fig.
7.1(c)). In some machines there is a difference in the angle α made by the roving
entering the grommet; this leads to differing false twist and tension between the front
and back rows. The result shows up as differences in yarn coming from one roving
bobbin and another; the differences can produce barré in fabrics. The solution is to
use extensions, which raise some grommets from B to A to give equal amounts of
false twist. Worn grommets can cause similar problems but they are occasional rather
than prevalent.
7.1.4 The roving process
Roving usually has a count of about 1 or 2 hank/lb irrespective of whether it is in
cotton or worsted processing. The count of roving is often quoted as ‘hank roving’.2
2 See Appendices 1 and 2 for definitions and calculations relating to Ne or Nw.
Short-staple spinning
173
Since this roving has then to be drafted to make yarn, and this drafting involves the
sliding of fibers over each other, it is not practical to insert a high level of twist in the
roving. Twist multiples (TM) of less than 1.0 are typical, but precise values depend
on the staple length, fineness (or denier), and type of fiber. Data given by Klein [2]
suggest that the TM for cotton is given by the regression TM = 1.785 – (0.46 × fiber
length in inches). Cottons need a higher twist than synthetics; coarse fibers need
more twist than finer ones. In addition, lightweight rovings need more twist than
heavier ones. More roving twist and the use of lightweight rovings lead to less yarn
hairiness, but too high a roving twist impairs the drafting operation in ring spinning.
High roving twist is liable to cause defects in the yarn. There has to be a compromise
between having a high enough twist in the roving to give it sufficient strength to
withstand processing and having a low enough twist to permit proper drafting in ring
spinning.
Imperfections in drafting at this stage are likely to lead to thick and thin spots in
yarns produced in ring spinning. A thick spot (or ‘slub’) is more difficult to twist than
a thin one, and when a strand containing thick and thin spots is twisted, the twist
therefore concentrates in the thin spots. Irregular roving is weak at places and overtwisted at others; it is difficult to process. Such poor material can cause considerable
problems at the ring frame because of the so-called hard ends or tight spots. It
becomes more likely that an undrafted portion of the roving is drawn through the
drafting system of the ring frame creating a thick spot followed by a thin spot in the
yarn. The weak spots give so-called creel breaks. A rule of thumb is that the roving
should be just strong enough to remove it from the bobbin, manually, in a direction
parallel to the axis of the bobbin. Differences will be found in this attribute if the
roving is stored for any considerable time. The twist becomes ‘set’ and the layers
more firmly bedded, with the result that it is sometimes difficult to process such aged
material in the ring frame. High humidity in the workplace can produce a similar
effect.
A machine is incapable of shutting down immediately and there is always some
discharge of fiber when an end breaks. In consequence, the whole frame is normally
shut down automatically when an end does break. When an end breaks in roving, a
free end of the weak roving lashes the adjacent structure creating a veritable snowstorm
of fibers until the machine finally stops. In addition, there is always fly (airborne
fiber) from other machines. Fortunately, this is not normally a frequent occurrence.
Nevertheless, it still requires vigilance by the operator and the ability to shut down
quickly. Uneven or faulty sliver should be avoided. It goes without question that the
machine must be correctly set. The solution of increasing the roving twist to reduce
end-breakages of the roving is rarely acceptable. There are a few machine designs
emerging which break out the sliver supply when an end breaks.
Fly contains dirty fibers and aggregates into masses, which are often picked up by
the sliver or roving to give an undesirable thick spot. These can cause flaws in the
materials or end-breaks in subsequent processes. A snowstorm of fiber also causes
considerable masses of unwanted fiber to be deposited on adjacent rovings and
serious yarn faults can also result from this. To reduce this danger, it is now standard
practice to install traveling cleaners, which patrol a series of machines to blow fibers
from sensitive areas and pick up the material disturbed by the blower. The cleaners
consist of long, vertical, pendant tubes with air nozzles protruding at appropriate
heights and a suction nozzle, which sucks the displaced fibers from the floor. These
systems of tubes are moved along elevated tracks that pass along the fronts of several
174
Handbook of yarn production
machines. The fibers swept up are filtered from the air and might even be reworked.
Substantial lint collection systems with adequate suction are also needed. The lint
collection systems have to be maintained, which involves regularly cleaning (or replacing)
the filters, and removing the reusable fiber. The air for blowing is usually supplied by
a blower on the traveling cleaner itself.
7.1.5 Apron drafting in roving and ring spinning
Both roving and ring spinning usually have apron drafting. It is usual to angle the
drafting systems with respect to the vertical (Fig. 7.2) to improve access and to help
control the fibers in the twist triangle. The top rolls hinge upwards, out of the way,
when the system needs rethreading, cleaning, or maintenance. Roving is removed
from the drafting system at a fairly shallow angle to the horizontal and there is a little
wrap around the front roll. Fully formed yarn is nearly always taken from the drafting
system with a partial wrap of the fibers on the front roll, which causes moderate
pressure to be applied to at least a part of the twist triangle. This gives a measure of
control of the weak ribbon of fibers emerging from the front nip. Many end-breaks
occur in this zone and control is important.
There is a choice between short and long aprons (Fig. 7.2(a) and (b)). Long aprons
are less likely to choke and are easier to fit than short ones but the long ones have a
larger initial cost. Short aprons have better fiber control. To some extent the choice
is governed by the rate of fly production and the class of yarns being made. Correct
aprons and settings are needed to control the unevenness of the yarn since the greatest
added variance to the product is created at the ring frame. Setting assumes a different
meaning from that applied to a drawframe, and an example of it is given in Fig.
7.2(c).
The extent of the variance is related to the high drafts used. Also, the use of correct
hardness of the rubber cots on the top rolls is important. Referring to the whole
drafting system, the back rolls have to control more fibers than the front ones.
Consequently, heavier pressures have to be applied to control the fibers, and harder
rubber is used there. Typical hardnesses of the rubber are 80° to 85° Shore at the back
and 65° to 70° Shore at the front rolls. Rolls tend to harden in use and vigilance has
to be maintained to spot any loss of fiber control due to the hardening of the rubber.
The softer rubbers wear more rapidly than the harder ones and the hardest rubbers
consistent with fiber control should be used. Also, soft rubber cots tend to lap more
easily. When fibers first lap a roller, a surface is created that tends to collect more
fibers, and very quickly a dense wrapping of fibrous material forms. This build-up
can be so powerful that it forces the rolls of a drafting system apart and can even bend
the shafts. The condition is created when the fibers are longer than a certain percentage
of the roll circumference, or when the fibers or the surfaces are sticky. When rubbercoated rolls are used, a build up of static electricity can also start the undesirable
process of roll lapping. Maintenance of correct rh in the vicinity of the rolls minimizes
this problem. A discussion of the phenomenon of lapping is given in Section 8.2.1.
Experience has to be used to make the final choice, depending on local conditions.
Related to this, it might be seen that maintenance is important; a normal cycle
includes buffing the rolls at intervals varying between 150 and 200 production days.
This interval depends on the fibers being processed, roll weighting, draft, and rubber
hardness.
In ring spinning, the possibility of a lap-up is increased when an end breaks and
Short-staple spinning
175
the fiber stream is diverted into the pneumafil waste suction system. (A pneumafil
waste system collects fibers emerging from the front rolls of a ring frame at those
times when an end is broken; it is uneconomic to stop the whole frame, or even a
section of it, for a single break.) Schiffler [3] postulates that the wrap frequency
varies inversely with time, logarithmically with respect to the number of fiber ends
entering the waste suction, and inversely with the apron clearance. Schiffler calls a
lap-up a ‘wrap’ and he goes on to say that the wrap frequency (the number of wraps/
unit time), is normally a very small number. The apron clearance is defined as the
distance apart of the aprons at the nose. The state of the rolls and the extent of ageing
of the rolls affect the issue, as does any change in fiber, roll size, weightings, etc.
Periodically, farmers suffer infestations of aphids and other insects, which eventually
produce contaminants such as sugars on cotton; the fibers become sticky and difficult
to handle in processing. Fiber lapping then becomes a problem. The effects of sugar
reduce gradually as microbial action breaks down the substance, but insect excreta
take longer to break down and sidelining the affected bales is then no longer a
reasonable option. In man-made fiber production, the application of too high a fiber
finish level can produce a similar result. Carelessness in tending the machines can
sometimes result in deposits of oil or grease on the rolls; this too can produce these
effects.
7.1.6 The ring spinning machine
The ring spinning machine took about two human generations to replace the mule
but, by 1982, some 150 million spindles were installed throughout the world, of
which 80% were used for short-staple spinning [4]. The ring frame consists of a large
number of spindles. One traveler and spindle co-operate with a bobbin, to twist and
wind the yarn from a drafting system as shown in Fig. 7.3(a). This sub-system is
replicated several hundred-fold in a ring frame because of the low productivity of a
spindle. As with the roving frame, the bottom rolls are sometimes long cylinders,
extending over many spindles, articulated at intervals along the frame and connected
to gearing in the headstock. There are now some very long frames of about 1000
spindles/machine, and articulation is necessary to prevent trouble from changes in
floor height that might distort the whole frame and cause bearing problems. The
spindles are driven by one or more tapes (thin flat belts), which engage the whorls
(pulleys) that project from the bottom of the spindle. Slippage of the tapes can lead
to twist losses, which vary from spindle to spindle and which, in turn, give barré and
streaking problems when the yarn is assembled into finished fabric. Consequently, it
is desirable to carry out periodic checks with a stroboscope to find spindles that are
out of tolerance regarding synchronism. This is not a small task because the spinning
area is very large and sometimes covers acres of floor space.
The ring frame is normally fed with roving from a large bobbin, and delivers yarn
to a smaller one. Because of this, the roving bobbins in the creel have to be renewed
less frequently than the yarn bobbins (ring tubes). When spinning a coarse count, the
ring bobbins have to be doffed every few hours. This used to consume considerable
amounts of labor. These days, the doffing is carried out automatically or semiautomatically and this process is referred to as autodoffing. At the time of writing,
there is also some use of automatic or semi-automatic creeling in which the roving
bobbins are transported to the ring frame by a rail system and the empty bobbins are
automatically replaced by full ones as necessary. This further reduces the labor needed.
176
Handbook of yarn production
Pigtail guide forms
node for yarn balloon
Roving bobbin
Drafting unit
Aprons
Ring
Components of the
system are not in scale
with one another to
enhance illustration
Separator plates
Ring rail
Traveler
(a)
Ring tube
Yarn
Ring bobbins
X
Enlarged view
of pigtail guide
(b)
Ring
Traveler
Ring flange
Enlarged view of ring and traveler
(c)
Fig. 7.3 A typical ring frame position
Typical spindle speeds for new machines in 1995 were in a range around
18 000 r/min, but new materials have been introduced, which inhibit ring and traveler
wear and it now possible to raise the speeds to about 25 000 r/min. However, when
spinning yarns with abrasive fibers or fibers with poorly formulated finishes, or fine
yarns, speeds have to be reduced. Occasionally, two rovings are fed to each spindle
(‘double creeling’) to even out errors by doubling, but the draft ratio is thereby
increased, which also increases the errors generated by the drafting process. The
result is a trade-off between improvement in long-term errors and deterioration in
short-term errors. The practice is avoided by many, purely for economic reasons.
High levels of error are often obtained with high draft. However, with attention to
design detail and proper settings, short- and medium-length errors can be reduced to
very acceptable levels. Nevertheless, high draft also tends to increase long-term
errors and to produce yarn hairiness, unless condensers are used in the main draft
zone [5]. Unfortunately, the lengths of the long-term errors fall outside the ranges
Short-staple spinning
177
normally measured, and often these defects are not detected until the yarns are
assembled into fabric.
Twist seeks the thin spots because of the low stiffness there; this phenomenon can
give exaggerated defects in the fabric. With a good, even yarn there is little problem,
but with an uneven one there may be complaints.
7.1.7 The twisting phase in ring spinning
In ring spinning, the energy to drive the twisting mechanism is derived from the
bobbin, but the level of twist is controlled by the traveler. The traveler is a C-shaped
piece, which slides around a flange on a ring that is set into a ring rail as shown in Fig.
7.3(a). The rotating yarn balloons out, and it is necessary to use separator plates to
prevent the clash of yarns from neighboring spindles. (If the balloon gets large
enough to impinge on the separator plates, the yarn becomes more hairy and the
spinning efficiency is impaired.) The mass of the traveler has to be balanced against
the yarn linear density, and the so-called ‘traveler weight’ is an important factor in
determining the yarn tension. The yarn tension, in turn, is an important factor in
determining balloon size as well as the end-breakage rate.
Each revolution of the traveler inserts one turn of twist into the yarn. There is a
twist gradient across the pigtail guide (Fig. 7.3(b)), and some false twist caused by
the yarn rolling on the internal surface of the pigtail guide. Consequently, the twist
above the pigtail guide is a little less than might be expected. The bobbin rotates
faster than the traveler and the trailing yarn drags the traveler behind it (Fig. 7.3(c)).
The difference in speed causes the yarn to wind onto the constant speed bobbin. The
yarn winds at different diameters during the build of the package and this causes
slight variations in the twist insertion rate, but the differences even out over the
length of the yarn.
Productivity of a ring spindle is very low. For example, a spindle producing a
4 TM, 36s yarn at 18 000 r/min and 95% efficiency delivers only 0.039 lb/hr.
Consequently the machine has to contain many spindles to achieve economy of
tending and workspace. These are reasons for the ring rails, which hold many rings
and serve many spindles. As an aside, it might be mentioned that it is also a reason
for the bottom rolls of the drafting system to cover many spindles. It is an economical
arrangement that reduces the capital cost of spinning.
7.1.8 Package build
To build a yarn package, it is necessary to move the ring rail, which extends the
length of the machine and carries the rings. This rail oscillates over about 2 inches
(≈ 5 cm), in an asymmetrical pattern, during which the spindle rotates several hundred
revolutions on, say, the upstroke and a much smaller number on the downstroke or
vice versa according to the design of the machine. This short oscillation is called a
chase; the term serves to distinguish it from the creeping build motion (Fig. 7.4(a)).
The asymmetric building motion gives a wind structure that is stabilized by the
longer yarn spirals generated on the downstroke. Each cycle of the chase creates an
interlocked double conical layer of yarn on the top of the package (Fig. 7.4(b)) and
this is called a weft wind. The actual number of rotations of the spindle can be set by
changing the appropriate gearing. Superimposed on this motion is a slow lift, which
changes the rail height sufficiently to accommodate the conical layer of yarn just
178
Handbook of yarn production
Creeping
motion
Chase
Bobbin
rotates
Bottom
rail
Belt or tape
Weft wind
(b)
Top rail oscillates and moves
(a)
slowly upwards during the build
Movement of element
Pigtail guide
Balloon
control
ring
Ring &
traveler
(c)
Fig. 7.4
Time
Package building
laid. Many machines not only move the ring rail but also control the position of the
balloon control ring and pigtail guide (Fig. 7.4(c)). This helps to keep the yarn
tension variations within a closer band than otherwise would be the case. The tube
height is about 5 ring diameters plus about 0.2 inch (5 mm) and the total lift of the
ring rail is about 0.8 inch (≈ 20 mm) shorter than this. The balloon height is about 6
ring diameters and this is important because an excessively tall balloon induces a
high spinning tension and increases the chance that the top of the tube will interfere
with the yarn in the balloon. These circumstances are undesirable because of increased
end-breakage in one case and variable yarn hairiness in the other. Some travelers
create more yarn hairiness than others.
7.1.9 Ring and traveler
Travelers are shaped to accommodate the ring flanges and flowing yarn; they are
normally made from small lengths of wire of a variety of cross-sections. There have
Short-staple spinning
179
been developments over the years in the profile and size of the ring flange, as well as
in the corresponding traveler.3 These developments have been aimed at increasing
productivity or reducing end-breakage rates. Also, the traveler has been changed to
lower the point of contact of the yarn to lessen the tilting action that leads to jamming.
A sampling of typical flanges and traveler profiles is given in Fig. 7.5. Diagram (a)
shows a typical shape and (b) shows a design of ring that has two flanges to permit
reversal so as to provide a new wear surface when the first flange is worn; this
extends the life of the ring. The SU ring shown in (c) has a conical flange that
distributes the load from the traveler to lessen wear and give greater stability, but it
has a tendency to be difficult to piece. Piecing refers to the repair of an end-break,
which is discussed in Section 7.1.11. Special techniques have to be developed to deal
with the piecing problem. All rings should be uniformly smooth and properly centered,
otherwise once-per-revolution variations in yarn tension occur and the end-breakage
rate is increased. Diagram (d) shows a selection of traveler wire cross-sections.
Round cross-sections are used for wool and long-staple spinning, whereas flat and
half-round cross-sections are used for short staple. Flat cross-sections are often used
for cotton because such travelers help clean the yarn; they shave off projecting hairs,
which help lubricate the ring, but produce fiber build-ups on the traveler. The halfround cross-sections are frequently used in elliptical travelers. Elliptical shapes give
a lower center of gravity to the travelers, which reduce the tendency for them to tilt
in operation. Some travelers are made from special wire, which more readily transfers
heat from the sliding surfaces and permits high speed operation. Both rings and
travelers should be run in at low speed for a period before they are used in high speed
operation. In practice, it is rather onerous to run in travelers and many do not do it.
Normally, no oils can be used for lubrication of the traveler, otherwise there is a risk
of oil stains on the product. Crushed fiber debris is usually sufficient for lubrication;
sometimes non-liquid anti-friction surfaces are used. When lubrication fails, shortlived micro-welds form, which disrupt the smooth movement of the traveler. The
choice of traveler is conditioned by the type of ring flange used, as well as by the
intended product and the production speed. If the machine elements are improperly
Yarn
Yarn
Yarn
F
F
F
(a)
Ring cross-sections
Ring flanges at F
(b)
(c)
Traveler
crosssections
(d)
Fig. 7.5 Ring and traveler cross-sections
3 Rotating rings have been tried in attempts to overcome the sliding problem just discussed, but
although productivity increases of up to 40% have been cited [6], the capital cost of the ring is
increased, and the lack of simplicity has prevented the system gaining significant market penetration.
180
Handbook of yarn production
placed, or the sliding surface on the ring becomes damaged by rust or micro-welds,
the defect is reflected in a periodic variation in yarn tension.
An eccentric spindle also produces variations. Since the probability of an endbreak is greatest at the peaks of yarn tension, it is the maximum tensions that matter
rather than the average. Thus, reductions in the deviations from the normal are important
in minimizing the end-breakage rates. The efficient running of a mill depends, in part,
on ensuring that the rings are well maintained and that the correct travelers are used.
The foregoing implies that the traveler must be changed when the yarn count is
altered and sometimes when the fiber is changed. If the traveler is too heavy, or the
spindle too fast, the load between the tiny traveler and the ring flange becomes
sufficiently high to cause excessive wear or burns. Travelers wear and have to be
changed frequently. Consequently, not only must the correct traveler be chosen for
the job but also the traveler changes have to be scheduled on a regular basis. Burned
and worn travelers can fly off and eye protection is advisable in the ring room.
Attempts to run the traveler too fast not only cause the contact area of the traveler to
burn, but also cause the yarn to be damaged.
Speed is limited by the traveler, which, in short-staple spinning, is rarely lubricated
with oils. Even if the travelers are nickel plated or otherwise treated, they can only
slide up to between 100 and 150 ft/sec (30–46 m/sec) during their short working
lives. The precise speed depends on the ring diameter, smoothness of the ring surface,
yarn tension, and the traveler weight and design. The ring surface must not be polished
since a suitable micro-structure of the surface is needed. The life of a traveler is very
short and is measured in days of running time. The combination of the various forces
acting on a running traveler causes it to tilt and this affects performance. According
to Klein [2], normal running pressure between the traveler and ring is 35 N/mm2;
consequently the ring has to have a hard, smooth (but not polished) surface and the
traveler has to have a less hard running surface so that there is sacrificial wear on the
cheaper traveler. As mentioned, it is good practice to ‘run in’ new rings to produce a
viable surface. This entails running at reduced loads and speeds for some hours
before resuming normal operational conditions.
It can be calculated that the speed of winding on the ring frame is only a very small
percentage of the spindle speed, which is fortunate because the spindle speed is fixed
and it is the traveler that controls the twist insertion rate.4 Some calculations in this
regard are given in Appendices 1 and 2. As the bobbin diameter changes, so a very
small variation in twist occurs. When the bobbin diameter builds from (say) 1 inch
(≈ 25 mm) to 1.75 inch (≈ 44 mm), the traveler speed might vary from 15 950 to
15 900 r/min, a difference of about 0.3%, which is negligibly small as far as twist is
concerned.
The outside diameter of the yarn on the package must be less than the ring size.
The diameter of the empty tube onto which the yarn will be wound has a minimum
size of at least 45% of the ring size, otherwise excessive yarn tensions would be
generated. The density of the yarn is approximately fixed. As mentioned earlier, the
length of the package is limited to about 5 ring diameters. This limits the changes in
yarn tension caused by increases in balloon height. Since practically the whole bobbin
has to fit inside the balloon at some time or other, the length of the package has to be
limited and therefore the mass of yarn that can be stored on a ring tube is often
limited to just a few ounces (1 oz ≈ 28 g).
4 It is not correct to say that the traveler puts in the twist; the energy derives from the spindle.
Short-staple spinning
181
Forces acting on the yarn passing through the traveler, and forces acting on the
traveler, have to be in equilibrium. There is an important balancing mechanism provided
by the traveler. The forces acting can be classified as (a) those coming from the yarn,
(b) centrifugal force acting on the traveler, (c) frictional force acting on the traveler,
and (d) the reaction component between traveler and ring, acting normal to the ring.
The resultant of force components (a), (b), and (c) is balanced by the reaction force
(d). Part of the centrifugal force acting on the traveler is absorbed by the ring at the
point of the sliding contact, and part applies tension to the yarn entering and leaving
the traveler. The very important balance between these two parts is affected by the
angles that the yarn and traveler adopt under the given running conditions. The
balance changes with the ring rail position and so do the forces (and tensions) involved.
The tension averaged over, say, one second, changes with the geometry of the system.
There is a variation that more or less follows the movement of the ring rail. During
the chase, however, the tension is not in exact synchronization with the rail position
because the winding point on the bobbin lags behind the recent movement of the rail.
In the main, the yarn tensions are at a maximum when the rail is near the top of its
chase but near the beginning of the build. Some machines control the spindle speed
according to the position of the ring rail, to restrain the yarn tension [7].
If the yarn tension is too high, the probability of an end-break increases. The
practical way of adjusting the yarn tension in mill practice is to alter the traveler
weight and type according to the yarn being spun. Changes in traveler weight and
design affect yarn hairiness, but there is some dispute as to the extent [8]. The traveler
weight also affects the yarn package density. Too light a traveler can produce an
undesirably soft yarn package with poor yarn storage capability, but what is more
important, it can shift the operational condition to a zone of instability when a bobbin
is in the early stages of build. The balloon collapse associated with the instability
leads to end-breaks. The complex subject of ballooning is considered in Appendix 9.
The traveler weight is selected on the basis of the yarn count and traveler type.
Travelers are described commercially by numbers in both the direct and indirect
systems of counting; for example, they are often quoted in grains per 10 travelers but
they might be quoted in the number of travelers per unit mass. However, it is helpful
to use a traveler mass unit related to the linear density of the yarn instead. In that
case, recommended traveler weights range from 2.6 mg/tex for high yarn counts to
3 mg/tex for counts in the range of 20 to 30 tex. The values given are subject to
adjustment for the type of traveler in use and the spindle speed.
7.1.10 Spindle eccentricity
Lünenschloss et al. [9] showed that eccentric spindles can produce increases in
hairiness of the yarn as well as influence its strength and elongation. This was especially
true at high eccentricities. With a 20 tex, 65/35 polyester/cotton yarn, the hairiness
changed from about 1000/m to 1700/m at a spindle eccentricity of 2.5 mm. With
combed cotton, the change at the same eccentricity was from about 500/m to 700/m.
Apart from its effect on the yarn, the life of an eccentric spindle is shorter than that
of one that runs true; also, the noise level is worse. An eccentric spindle, or a displaced
guide or ring, can increase the end-breakage markedly because of the once-perrevolution cycle of tensions produced; this has important economic repercussions.
(Note: 407/yd ≈ 500/m, 640/yd ≈ 700/m, 914/yd ≈ 1000/m, 1550/yd ≈ 1700/m,
0.1 inch ≈ 2.5 mm, 20 tex ≈ 30 cotton hank/lb (Ne.)
182
Handbook of yarn production
7.1.11 Dealing with end-breaks
End-breaks cause a loss in production because the spindle produces no yarn after an
end-break until it is repaired. An operator may serve up to several thousand spindles
in modern plants and it is unavoidable that, on occasions, an end-break will cause a
delay of perhaps twenty minutes or even longer before a repair is made. If an endbreak were to occur every hour on every spindle, the production efficiency would be
terrible; even if it occurred once a day, there would still be a significant production
loss. An end-break has to be a rare event if reasonable production efficiencies are to
be maintained.
When an end breaks, normally the textile fiber keeps flowing and is sucked away
by a pneumafil or waste fiber suction system. To manually repair the breakage, the
operator has to retrieve the end from the bobbin, thread it through the traveler,
balloon control ring (if one is fitted), and the pigtail guide before inserting it into the
nip of the front drafting roll. With an experienced operator, this takes only a second
or so, but the time spent in patrolling to find the end-break is quite another matter.
Machines, or attachments to the ring spinning machine, can simulate the action, but
the capital cost is high.
It is also possible to interrupt the roving supply (known as a ‘roving stop’ system)
to prevent wastage and choking. The complexity of the roving stop results in an initial
cost that amounts to a significant sum because of the thousands of spindles involved,
but it does reduce the fiber loss significantly and it improves yarn quality as a
consequence. Because of the capital costs involved with the roving stop system, most
users prefer the conventional method but it involves the expense of dealing with
about a 2% fiber loss.
Waste fiber can be recycled, but only with care because it does not spin well. It has
to be mixed and diluted with virgin fiber. In spinning, there is always some fiber loss
from the twist triangle zone and suction is always required to remove the waste.
When an end breaks, the amount of waste increases. Over all the spinning frames,
there is a level of waste that is dependent on the mean end-breakage rate and beyond
a certain level it becomes difficult to absorb the wastage without deterioration of the
mill performance or the quality of the product.
7.1.12 Ring frame limitations
As already mentioned, the ring size is limited on the large side because of traveler
burns at normal production speeds, but, on the other hand, the ring cannot be too
small, otherwise the bobbins would hold so little yarn that the cost of changing
bobbins would be prohibitive. One of the limitations of the ring frame is the traveler
speed. With non-rotating steel rings and steel travelers, the linear speed is limited to
about 100 ft/sec (≈ 30 m/sec). The limit arises because the poorly lubricated traveler
makes micro-welds with the surface of the ring, which are immediately broken as the
traveler goes on its way. This creates incremental damage to the surface of the ring,
which still endures over the life of many, many travelers. However, the roughening of
the ring surface also progressively shortens the life of travelers. One solution is to use
ceramic materials for the rings to minimize the damage.
For a frame to run at a higher speed, the ring diameter has to be smaller and clearly
there is a limit to how small the ring can be. The volume of yarn on the package is
less than ρ( π /4)( D12 – D22 ) L, where D1 is the outside diameter of the package which
must be less then Dr, the ring inside diameter, D2 is the diameter of the bobbin, and
Short-staple spinning
183
L is the length of the bobbin covered with yarn. D1 must come as close to Dr as
practicable and the corresponding mass needs to be as large as the limitations in ring
geometry will permit. Currently the ring sizes for short-staple spinning vary from 1.4
to 2.1 inches (≈ 36 to 53 mm) inside diameter and they use several sizes of flange
varying from 0.125 to 0.16 inches (≈ 3.2 to 4.1 mm). Smaller rings mean not only that
the mass of yarn that can be stored is reduced but that the frames have to be doffed
more frequently. With automatic doffing and good splicing this is of less importance
than formerly. However, doffing still takes time from production and the smaller the
ring, the lower the spinning efficiency (but only incrementally so). An offsetting
factor can be that smaller rings mounted on a machine that is designed to take them,
reduces the capital cost.
Another limit is reached when spinning medium to fine counts of highly twisted
cotton and blend yarns, in which case the speed has to be reduced to suit the prevailing
conditions. The heat dissipation at the traveler is approximately proportional to the
third power of spindle speed and the traveler has to be able to dissipate the heat
generated by conduction or radiation. When the lubrication breaks down, the safe
temperature limits of either the yarn or traveler might be exceeded. In the one case
there are molecular changes to the polymer structure, or even surface melts of the
fiber. In the other case, the metal structure changes; the metal changes color and
eventually breaks. One solution is to run with a lighter traveler but if this is carried
too far, the balloon collapses and there are ends down as a result. Of course the speed
can always be reduced but that begs the question because we look to higher speeds to
improve economic performance.
Yarn tensions are also responsible for a good portion of the excess end-breaks in
spinning and this is why attention is focused by some on the ballooning mechanics
(see Appendix 9). Figure 7.6 gives an example of tension variations in which the
black line is the running average over a 0.2 second interval and the light gray points
are the variations from this running mean [10]. The running average shows the
variation caused by movements of the ring rail, with some superimposed variations
due to changes in linear density. When the tension variations exceed a level indicated
by xx, the tensions become high enough to pose a risk of breakage. The probability
of an end-break is determined by the statistical distributions of yarn strength and
tension. The height of the line xx is influenced by the strength of the weakest links in
the yarn being spun. The points at risk are shown as small black crosses and the
number of these should be rare if there is to be reasonable spinning efficiency. Klein
[2] reports that the majority of end-breaks occur as the ring rail approaches its
topmost position in the chase, rather than on the downstroke.
7.1.13 Mill balance
Returning to the question of productivity, let us establish a range. Compare the
productivities per spindle when making 81s and 9s yarn, both with a TM of 4.0. The
first is a fine yarn and the second is a coarse one.
Assuming the spindle speed to be 18 000 r/min and the efficiency to be 95%, the
productivities are 0.012 and 0.314 lb/hr respectively (using the formulae in Appendix
A1.5.1). Considering that a mill might produce, say, 20 tons/day (1867 lb/hr), at least
155 000 and 6 000 spindles, respectively, would be required. The way in which
productivities alter with average count leads to an operational difficulty. Suppose the
mill had been set up to produce 36s. The number of spindles required would have
184
Handbook of yarn production
When the specific tension is lower than
the breaking stress (xx), then the
chance of an end-break is low.
When the specific tension is higher than
the breaking stress (xx), then the chance
of an end-break is high.
40
x
x
Specific tension (g/tex)
30
20
Running average of
specific tension
10
0
0
10
Time (seconds)
Fig. 7.6
20
Yarn tension in spinning
been about 48 000. If a heavier yarn such as the 9s were substituted, there would be
42 000 excess spindles, whereas, if the very fine 81s were substituted, there would be
a shortfall of 107 000. In this exaggerated first case, over 87% of the ring frames
would be shut down with an accompanying loss of employment for the operators. In
the second case, also exaggerated, the mill would be capable of supplying only 30%
of the requirement from production. The figures are exaggerated to make the point;
normally the average count is kept reasonably near the design value. It is very expensive
to change the so-called balance of the mill, once it has been set up.
7.1.14 Automation in ring spinning
At this stage, automation from sliver through to yarn will be discussed. Sliver handling
is automated to different degrees according to circumstances. Nearly all modern
cards and drawframes have automatic doffing. Some mills use automated guided
vehicles (AGVs) to marshal the cans in the creels of the following machines. In rotor
spinning, this segment of the costs plays a larger proportionate role than it does in
ring spinning.
With roving, the usual solution is to use an overhead rail system with the bobbins
suspended from carriers. In many of the systems, the transport system carries the
bobbins from station to station, passing occupied positions and making exchanges for
an empty bobbin where a full one is needed. There is a risk with this system that a few
bobbins might circulate for a long time before they find a home. Periodic checks for
such ‘joy riders’ prevent difficulties. With roving, it is necessary to control the loose
Short-staple spinning
185
ends during transport and this is normally accomplished by making a wrap of the end
in a secure place before doffing. The use of AGVs is also possible.
Mechanical piecers have been introduced from time to time. They work well but
they have not become established. There is a feeling that there is still a need for
patrolling operators to clean, check on performance, and perform other duties.
Automatic ring frame doffing has been widely accepted and the most common
system involves rails that reach from end to end of the frame. These rails are designed
to carry the full bobbins during the doff, and the empty ones during the replenishment
phase. The doffer rail carries apertures for each spindle and each spindle is equipped
with a grasping device. The grasping device is often an inflatable cuff which fits over
the bobbin and grasps it. The purpose is to lift the full bobbin from the spindle
without damaging the yarn. Two series of pegs are mounted on a belt running the
length of the machine. One series of pegs carries empty bobbins which have been
mounted before the start of the doffing sequence. The ring frame is stopped automatically
when the bobbins are full, then: (a) the ring rail is lifted clear after the ends of yarn
have been trapped at the base of the spindle; (b) the doffing rail is dropped over the
full bobbins; (c) the grasping devices are activated and the rail is used to lift the full
bobbins from the spindles; (d) the full bobbins are deposited on the vacant pegs on
the belt just mentioned; (e) the doffing rail then picks up the empty bobbins from the
belt; and (f) the rail deposits these empty bobbins on the empty spindles. On start-up,
the yarns should still be threaded through the travelers and the rotating bobbins
should catch the yarn and start spinning automatically. In practice, a few ends fail to
catch and have to be pieced manually. Thereafter, the belt moves towards the end of
the spinning machine and the full bobbins are either removed or continue on to the
winder. When the bobbins are transported directly from the autodoffer to the winder
without human intervention, it is known as ‘linked spinning’.
Clearly, some sort of quality control is needed because deformed or improperly
filled bobbins are unlikely to unwind properly and it is best to discard them. The
defects would be difficult to trace if not caught before winding. Consequently, the
bobbins are gauged at the exit from the spinning machine and faulty ones are rejected.
Looking to the future, monitors could be fitted to keep track of the number of endbreakages and other performance data. The cost of these monitors is high and few
mills are willing to invest in them until there is a more assured way of translating the
large volumes of data they provide into an effective control system, capable of yielding
an economic gain. So far, the complex factors and interactions in the process are not
well enough understood to permit accurate prediction of the outcomes.
7.2
Open-end spinning
7.2.1 Basic principles
The basis of open-end (OE) spinning is that fibers are added to an ‘open-end’ of a
yarn, as indicated in Section 3.4.1. Twist applied to the newly added fibers converts
them into yarn and the new elements of yarn are continuously removed from the
twisting zone. The theoretical advantages of such a system are: (a) it is easier to rotate
the small open-end of the yarn than it is to rotate a whole yarn package as in the case
of ring spinning; and (b) the twisting and winding can be separated. The first point
implies a tremendous potential for increased productivity and the second point means
186
Handbook of yarn production
that the package size is limited only by the design of the winder, which leads to
greatly reduced handling costs.
To produce an open-end, it is necessary to use a very high draft so that the fiber
flow is reduced to just a few fibers in the cross-section. This prevents twist from
running back into the fiber supply to produce false twist, which would defeat the
object of the exercise. The technology has developed to the point where OE machines
are fed with sliver and this eliminates the roving frame. A sliver might have (say)
20 000 fibers in the cross-section, and, if the fiber flux just before the open-end is as
low as (say) 2, the initial draft would then be 10 000:1. It is not possible to use roller
drafting alone to produce this sort of draft at the speeds required. Rather, it is more
normal to use a toothed roller, which acts in much the same way as a licker-in in a
card. The emerging yarn might have several hundred fibers in the cross-section and
thus there has to be a condensation of fibers leaving the open-end. It follows that the
essential phases in the spinning operations are as shown in bold font:
1
2
3
4
5
6
7
8
Drafting.
Fiber transport.
Fiber alignment etc.
Cleaning (if necessary).
Fiber condensation.
Twisting.
Yarn removal.
Winding.
In practice there have to be some intermediate phases as well and these are also listed.
7.2.2 Open-end systems
There are many embodiments of the basic idea of OE spinning and, although only one
(rotor spinning) has taken a large market share, it is worthwhile to mention briefly a
number of the other contenders.
The early mechanical systems of the nineteenth and twentieth centuries were too
cumbersome to work at high speeds. Generally, it was the invention of the mechanical/
pneumatic systems that led to workable prototypes, which had a chance of commercial
development. An early variety of OE spinning used an air vortex device. Another
variety was friction spinning, but although this reached industrial production, the
fine yarn version did not develop fully due to lack of yarn strength and other problems.
Friction spinning for coarse yarns and core spinning did become established for a
segment of the market. One successful form of friction spinning was the DREF
machine in which fibers are ‘rolled’ into yarn by a pair of condenser rollers as
sketched in Fig. 7.7.
Rotor spinning itself is now well established. In the early days, a rotor was easy to
piece up at the, then, low speeds, but it produced a yarn that was some 20% weaker
than ring yarn. Because the yarn was acceptable for some uses, and because of the
economy of operation, rotor spinning established a foothold despite these difficulties.
A further disincentive was the high capital cost component of yarn produced; it could
only be made to pay its way if low count yarns were made. Thus, there was pressure
to increase the count range spinnable, by increasing the operating speed. The first
commercial machines operated at 30 000 r/min, but, by the mid-1990s, speeds of
120 000 r/min were possible. Productivity raced ahead of capital costs to the point
Short-staple spinning
187
Input fiber
Airflow
Feed roll
Combing roll
Output yarn
Perforated suction rolls
Fig. 7.7
Friction spinning
that the break-even count, discussed so much in the 1970s, was no longer an issue.
That is not to say that the capital cost did not go up. For example, it was found that
manual piecing was impractical at the high speeds now established; automatic piecers
had to be designed and manufactured. Such further developments have a considerable
cost that has to be passed on to the machine buyer in the form of higher absolute
capital cost. Nevertheless, the capital cost of the machine/lb of yarn produced was
reduced.
7.2.3 Rotor spinning
Rotor-type OE spinning was first used commercially in 1969 [11], but by 1995 it was
a major part of the US yarn spinning capacity and had found widespread use elsewhere.
At the beginning of the development, rotor spinning was offered for both short- and
long-staple spinning but long-staple rotor spinning did not become established. Further
discussion will be limited to short-staple rotor-type OE spinning.
7.2.4 Drafting in rotor spinning
It is possible to combine a rotor with a combing roll drafting system to make a
spinning machine, so let us consider the drafting system before explaining the functions
of the rotor. The combing roll drafts fibers in much the same way as does a licker-in
of a card. Fibers are detached from a beard presented by a feed roll, as shown in Fig.
7.8(a); the detached fibers pass into the rotor at much higher speeds than the advance
of the beard and thus there is a drafting action. To sustain the high exit fiber speeds,
an adequate airstream is needed to carry the fiber forward in a sufficiently opened
state to prevent twist from running back to the drafting system. If one required the
ideal exit fiber flux of unity, the draft ratio in the combing roll would be equal to the
number of fibers in the cross-section of the sliver feed and that could number tens of
thousands. In practical terms, the fiber flux is usually greater than unity, but the value
cannot be very high or the system would not work; draft ratios at the combing roll are
usually measured in thousands. The fiber flow is later condensed inside the rotor as
188
Handbook of yarn production
Rotor
Fiber to rotor
Toothed
combing
roll
Feed roll
Cleaning
edge
Pressure
Combing roll
Sliver
input
Trash
out
Sliver
Feed roll
Combing edge
Feed plate
Bearing
Fiber beard
Fiber to rotor
Tape
Reaction
(b)
(a)
Fig. 7.8
Rotor spinning drafting system
a precursor to yarn. The overall draft ratio is that calculated from the ratio of linear
densities of input sliver and output yarn, measured in compatible units; it is the
mathematical product of the draft at the combing roll and the fractional draft (i.e.
condensation) in the rotor.
Various feed systems are possible, but the one that has become established is the
roll and table type shown in Fig. 7.8(b). Sliver is gripped between the slowly moving
feed roll and a stationary plate; the passageway reduces in size to increase the pressure
at the point where the fiber beard is formed. The action of the teeth of the combing
roll on the fringe removes fibers; these fibers are then discharged into the rotor where
the precursor to the yarn is formed.
The combing roll can be clothed with needles or saw-toothed wire but the latter
has established itself firmly in the market. Hunter [12] notes in his survey that many
find the pins superior to wire for opening capability and wear; he cites up to 60 000
hours’ life when spinning cotton. However, hardened and ground card wire with a
surface of diamond dust embedded in nickel has been developed since then [13].
Other sharp edges prone to damage are similarly treated with wear-resistant materials.
Siersch [14] found that there were advantages to spirally wound saw-toothed clothing
because the helix angle of the wire causes successive teeth to penetrate the beard in
positions that move across the beard as the roll rotates, and this gives a good distribution
of fiber separation. He found that saw-toothed wire gave lower CVs in the yarn than
did comparable needle arrangements. Wire-wound clothing is recommended for cotton
and cotton blends running at high throughputs, whereas pinned combing rolls are
recommended for fragile fibers such as acrylics and rayon running at moderate
output rates. The force acting on the fiber beard increases ever more rapidly as fiber
length is increased; beyond a certain level, fiber breakage then becomes a problem.
Thus, the device is best suited for short-staple fiber. The fibers can disengage the
teeth soon after leaving the combing zone and travel along the inside periphery of the
combing roll housing at a velocity lower than the surface speed of the combing roll
[15] (the housing is shaped to allow for the fiber flow). Friction between the housing
Short-staple spinning
189
and the fiber slows the fiber until it reaches the duct that carries it to the rotor. An
airstream is generated by the suction in the rotor, which accelerates the fibers and
straightens them somewhat. Lünenschloss [16] states that the fiber velocity at the exit
of the transfer duct is determined primarily by the negative air pressure inside the
rotor (i.e. suction). Without an adequate airstream, the fibers relax, become disorientated,
and tend to clump together.
It is desirable for the feed material being presented to the combing roll to be as free
from hooks and tangles as possible, in order to reduce fiber damage. However, it is
difficult to keep the entering fibers aligned as they are manipulated by the combing
roll. Stalder [17] has shown photographs illustrating how a single fiber can lie across
several rows of combing roll teeth before it is carried away into the rotor. Lawrence
and Chen [15] showed that short fibers were removed, and that longer fibers abraded
the edge of the feed channel.
Combing roll speed affects the yarn hairiness [18, 19] and can affect nep production.
Too high an overall draft ratio can give high end-breakage rates but good trash
removal; values in the order of magnitude of 100 are usually satisfactory. For cotton,
the sliver weight depends on the sliver preparation and the yarn count, but common
sliver weights vary between 50 and 70 grains/yd (≈ 3.5 to 5 g/m); for man-made
fibers, the slivers should be about 25% lighter than with these.
Fiber orientation in the sliver feed channel is important in terms of yarn tenacity
and evenness. For coarse yarns in which yarn strength is not very important, it is
sometimes possible to use card sliver as the input to the open-end spinning machine.
For such carded yarn, performance is enhanced if the sliver is drawn in the carding
machine. Otherwise it is normal to use drawn sliver. Occasionally, combed sliver is
used. Another way to lessen fiber damage is to reduce the number of fibers in the
fringe being combed by the combing roll, and this is achieved by using a lighter
sliver, which adds to the cost of production. This can be offset against the advantages
of rotor spinning. Fine combed sliver is difficult to handle.
7.2.5 Combing roll performance
The cleanliness of cotton feedstock depends upon the degree of cleaning at the gin,
opening, and carding. Cotton fiber ends tend to be damaged by the combing roll and
at higher combing roll speeds there is fiber breakage. The combing roll speed is an
important factor in this respect and this speed also affects the cleaning capability of
the combing roll system (Fig. 7.9). A cleaning edge separates most of the trash before
it can enter the rotor but it is difficult to remove all the dust (but often, when spinning
man-made fibers, no cleaning edge is used). Separated trash passes to a ‘dirt box’,
which is emptied periodically. Even when spinning man-made fibers, fiber breakage
Non-lint
(% of sample)
30
Rotor
20
10
Dirt box
0
4000
6000
8000
Comber roll speed (r/min)
Fig. 7.9
Trash extraction
190
Handbook of yarn production
and the deposition of debris in the rotor is not eliminated. Polyester fiber finish, fiber
debris, and oligomers are stripped from the fiber in passing through the combing roll
and are then deposited in the rotor. Naturally, weak fibers are liable to breakage, but
also damaged fibers have a tendency to pill. An optimum comber roll speed needs to
be established for each product to give the best compromise between productivity and
product quality. The clearance between the comber roll flanks and the casing should
be strictly controlled [20], otherwise airflow can cause fibers to become trapped and
jam the roll.
Unfortunately, dust and particles tend to be deposited unevenly, particularly if
large particles enter the rotor. Uneven deposits cause periodic unevenness in the yarn
and produce unacceptable moiré patterning in the fabric made therefrom. As trash
builds up, the quality of the yarn suffers and drifts lower as time elapses; when the
rotor is cleaned there is a sharp improvement. Changes in evenness, appearance, and
strength are caused by these cycles of rotor fouling and cleaning. They pose significant
quality problems unless controlled. This circumstance has sparked the development
of automatic rotor cleaning, which, in turn, requires automatic piecing to be effective.
Cotton fibers can be worked at high speeds but man-made fibers are usually
worked at lower speeds. The shape of the comber roll teeth is important because an
aggressively forward-raked tooth, such as is used with cotton, can be seriously eroded
by wear. This is especially true if it is used with dusty cotton or certain man-made
fibers. Dusty cotton is sometimes found after particularly dry growing seasons. Improper
or damaged wire can also produce neps, which can give serious quality control
problems. Also, abrasive materials from fiber finishes, fiber debris, and silica dust
from dusty cotton can cause excessive wear. Sharp edges (such as those at A and B
in the line of fiber flow, Fig. 7.10(a)) and combing roll teeth (Fig. 7.10(b)) are subject
to damage. It is desirable to avoid such finishes and fibers but, where this is not
possible, the combing roll speeds, the sliver weight, and the rotor speed should be
kept low.
Yarn
Rotor
Navel
Wear on tips of teeth
Doffing
tube
F
Fiber supply
channel
Cleaning
edge
B
A
F
Nicks on sharp edges
Cleaning
aperture
Comber roll housing
(a)
Fig. 7.10
Comber roll housing
Relative fiber motion shown by arrows marked F
(b)
Comber roll wear
Short-staple spinning
191
Lawrence and Chen [15] describe how short fibers tend to be thrown out of the
trash escape aperture while longer fibers are retained for a little more time on the
combing roll wire. The suction applied to the rotor removes the fibers from the
combing roll teeth but the fibers drag over edge B as they pass into the rotor. This
causes wear; both edges A and B have to be reinforced with especially hard material.
The fibers become damaged at these points also, with the result that dust and nep are
sucked into the rotor. Not surprisingly, the extent of trash separated by the combing
roll varies with the state of cleanliness of the sliver supplied as well as with the speed
and design of the roll. Rotor machines do not necessarily have such cleaning devices.
Machines without trash removal are more suitable for spinning man-made fibers, but
there is still a build-up of debris in the rotor due to the accumulations of fiber finish
and fiber debris. In all cases, the debris in the rotor causes a deterioration in the yarn
characteristics so it is important that the rotors be cleaned periodically. The length of
the period between cleaning depends on the type of fiber being spun as well as the
speed and type of spinning machine concerned. The combing roll speed is particularly
important in this respect. A 3 inch (≈ 76 mm) combing roll normally runs between
5000 and 8000 r/min; the higher the speed, the more fiber damage ensues and the
more debris is deposited in the rotor. However, the higher combing roll speeds tend
to give better trash separation. Trash build-up causes a gradual deterioration in yarn
quality. In particular, yarn hairiness, nep, unevenness, and other fault rates increase.
Barella [19] and many others found that regular cleaning of the rotor is required to
preserve yarn quality.
Choking of the feed mechanism must be avoided. If too heavy a sliver is fed, or if
the end inserted into the feed roll nip is doubled, it might overload the combing roll,
causing it to jam. During running, badly stored sliver in the feed can cause a loop of
sliver to be lifted from the can and to be fed into the drafting system. This produces
a similar unfortunate effect. If such choking is allowed to persist and the machine
continues to run for a long time, the whorl becomes overheated and so does the
surface of the drive belt. The result is that the contact surface of the belt becomes
glazed, causing slippage in all the combing rolls in the set. Such slippage might be
uneven in time and from combing roll to combing roll. This has an adverse effect on
yarn quality. Over long periods of time, overheating the combing roll assembly can
cause the grease in the bearings to harden and add to the difficulties by causing some
combing rolls to slip. Therefore, associated with choking is the possibility of bearing
damage. Combing roll bearings with race tracks indented by the balls have been
detected. Such damage increases the power requirements and increases the noise
level in an operating machine. An increase in power demanded by a comber roll
assembly increases the risk of slippage and of deterioration in yarn quality.
Most combing rolls are driven from a single tape or belt. Auxiliary pulleys, spaced
along the belt, control the path taken by it and apply the necessary force between it
and the whorls at the bottom of each drive shaft. These arrangements are discussed
in Section 7.2.8.
Increases in combing roll speed are often associated with increased output. However,
high combing roll speed and point populations can increase fiber breakage and lead
to increased end-breakages in spinning, as well as a reduction in yarn strength. Some
experimentation by the user is called for, to find the best compromise between machine
productivity and quality for the particular product. An exceedingly high machine
productivity does not necessarily produce the best financial return. Some trend curves
for combing roll performance when spinning cotton are shown in Fig. 7.11.
Handbook of yarn production
Thin
Thick
Nep
Ends down/h
Yarn faults
Yarn evenness
Yarn tenacity
192
Tooth angle
Fig. 7.11
Comber roll speed
Trend curves for a comber roll performance
7.2.6 Fiber flow into the rotor
Most machines will spin cotton or short-staple synthetic fibers. The trash in cotton is
deposited in the rotor at a fairly rapid pace unless it is removed before entry into the
rotor and it is difficult to remove all this trash by conventional means. Consequently
some rotor machines have a cleaning edge or cleaning aperture built into the combing
roll housing, as shown in Fig. 7.10. However, it is still of very great importance to
clean the fibers well in the opening and carding operations, otherwise the deposition
of dust in the rotor becomes a very severe problem with important economic
consequences.
Fibers flow from the combing roll through a fiber transport channel and are assembled
in the rotor, where yarn is formed. Fibers must be completely removed from the
combing roll and be transported to the rotor without being crumpled, disoriented, or
clumped together. This means that an airflow velocity exceeding that of the surface
of the combing roll must be used. To get such an airflow, it is necessary to run the
inside of the rotor at a vacuum of several inches of water.5 Today, the practice is to use
an external fan. The inside of the rotor gets hot and the fan fulfills the purpose not
only of inhaling the fibers into the rotor but also of removing hot dusty air from it.
The high temperature reduces the rh of the air inside the rotor and levels as low as
20% have been recorded. It is necessary to properly condition the input sliver. The
ratio of air speed to that of the surface speed of the combing roll should be 1.5 to 2.
(Some authorities suggest 1.5–4.) This is to ensure that the fibers are doffed properly;
too high an airflow can increase fiber waste.
5 An inch of water denotes a difference of air pressure equal to about 1/400 of an atmosphere.
Short-staple spinning
193
The disposition of the transfer channel or duct with respect to the combing roll is
important. Lawrence and Chen [15] showed that short fibers are thrown from the
combing roll at the transfer duct opening, and travel along the tube with what was the
leading end still leading. Longer fibers are dragged by the combing roll teeth over the
edge between the combing roll and the transfer duct, and then they are aspirated into
the duct with their erstwhile trailing end now leading. The hooked fibers entering the
transfer duct have a shorter fiber extent than if they were straight and, if such fibers
remain hooked or convoluted in the yarn produced, they reduce the strength of the
yarn. Thus it is desirable that some fiber straightening mechanism be employed. For
this reason it is common practice to use a tapered duct that accelerates the flow of air
and fiber as it approaches the duct exit. This tends to straighten the fibers; however,
observation of photographs shows that the acceleration is insufficient to remove all
the hooks and other fiber deformations in the transit stage. Fortunately, provided the
suction is not too strong, the surface of the rotor that first contacts the fibers emerging
from the transfer duct will be moving faster than the fibers. This sliding contact tends
to straighten the fibers [21] although they are rarely, if ever, completely straightened
and parallel at this point. The speed of a fiber as it enters the rotor should be about
80% of that of the metal surface on which it lands; the transitional draft at that point
should be between 1.25 and 1.4.
Fibers entering the rotor are deposited on the internal sliding wall (Fig. 7.12(a))
and move on this surface to the rotor groove, where they collect to make the fiber
ring. Figure 7.12(b) shows the sliding path,6 which is approximately fixed in space
with the rotor moving relative to it. Except in zones where the fibers sliding up the
inside of the conical portion of the rotor interfere with the outgoing yarn, fibers are
usually laid in the rotor groove in an amazingly parallel, straightened fashion. The
sliding wall is part of the transit system; surfaces must be well designed and they
must be kept clean. Fortunately, the movement of the fiber cleans the surface.
It must be realized that the yarn rotates at high speed with respect to the sides of
the rotor groove; centrifugal force presses the rotating yarn into the groove. Furthermore,
the wedge action of the acute vee of the groove causes the centrifugal force to be
magnified. Any abrasive particles that might be present then heighten the ‘lapping’ or
abrasive action of the rotating yarn. For this reason, the inside surfaces of the rotor
are treated to resist wear. Steel alloy rotors have been developed and it is now standard
practice to diamond coat the surfaces. The condition of the inside of the rotors is of
great importance.
Yarn is removed through a doffing tube.
7.2.7 Twist in rotor spinning
Real twist is applied to the yarn by the motion of the rotor acting on the yarn arm that
passes from the rotor groove to the yarn withdrawal point inside the rotor. Each
revolution of the rotor causes about one turn of twist to be inserted into the yarn, and
1/τ inches of yarn are removed (τ is the twist in tpi). There can be movement of fibers
with respect to the metal of the rotor during twisting. This is because the fibers are
not firmly held by any discrete nip at this point. Twist usually runs back along the
rotor groove and some fibers are laid onto an already twisted core of fibers. This
6 For economy of line in a complicated diagram, the picture shows fibers aligned along the sliding
path, but this is not necessarily true as they can move crabwise along the path.
194
Handbook of yarn production
Yarn
Yarn
withdrawal
tube
Rotor groove
Fibers
Rotor
Yarn
Fibers
Navel
Fibers slide on wall of rotor
during entry
Fiber sliding path
(a)
(b)
Yarn in rotor groove
N.B. Some twist runs into the rotor groove
Yarn arm rotates
False twist + real
twist in yarn arm
Real twist in
departing yarn
Navel
Enlargement of navel
(c)
Fig. 7.12 Fiber and yarn in the rotor
affects the yarn structure. The center of the end of the yarn withdrawal tube is fitted
with a non-rotating navel7 through which the departing yarn flows, as shown in Fig.
7.12(c). Sometimes a separator plate is introduced to prevent the premature capture
of the incoming fiber by the outgoing yarn. This makes a less desirable transport
system because of the complexity of the passageway, but in separating the incoming
fibers and outgoing yarn it fulfills a useful function.
The yarn entering the navel rolls on the inside surface. This rolling action produces
a false twist in the section of yarn inside the rotor. The false twist is in addition to any
real twist created by the movement of the rotor. Twist is trapped in the running yarn
between the point of twist application and the nearest upstream twist trap. In the
present case, the point of twist application is at the navel and the twist trap is on the
collecting surface of the rotor. The flare radius of the navel affects the false twist
generated, as shown in Table 7.1. Spinning performance and yarn character depend
on the yarn twist inside the rotor (which is false twist dependent) rather than just on
the apparent twist in the yarn delivered. Often, navels are grooved, as shown, to
increase the false twist, but as speeds rise there is less need for this. Also grooved
navels tend to make the yarn weaker as well as more bulky, neppy, and hairy, particularly
7 Experiments have been made with rotating navels, but they have not gained acceptance in the
market.
Short-staple spinning
Table 7.1
195
Effect of navel radius
Yarn linear density
59 tex
fiber
Navel radius (inches)
Yarn tenacity (gf/tex)
Yarn elongation (%)
CV of linear density (%)
0.06
12
18.0
6.2
25 tex
polyester fiber
0.2
21
17.6
9.1
0.06
22
13.9
7.1
0.2
11
11.0
8.1
at high rotor speeds. The grooves cause the yarn to bounce off the surface of the navel
for very brief periods of time. Yarn tensions measured inside the rotor are very close
to the theoretical figure given by the formula ω 2 rr2 n but there are pulses due to the
yarn riding over the grooves, if there are not too many of them. Unpublished work at
NC State University showed that the number of pulses rose with the number of
grooves until four grooves were cut. Increasing the number from four to eight gave
only four pulses and this was interpreted to mean that the yarn jumped over alternate
grooves.
False twist is also affected by how close the front surface of the navel is set towards
the flat inner surface of the rotor. Local shear in the air is produced by the rotor wall
moving close to the stationary navel. This shear can produce a small amount of false
twist in the yarn. Enlarged portions of yarn can interact with this space if the gap is
set too narrowly. Gages are used to set the distance accurately.
There will be differences in the coefficients of friction of the navel surface and the
yarn; also the navels wear. It has become common to use ceramic navels because of
their long lives but there can be unacceptable differences in the surfaces. As the navel
varies, it causes the nature of the yarn, and the efficiency of the operation, to vary.
Thus, it becomes important to make sure that all navels used in a given lot of yarn are
similar, or operational and barré problems will result.
The range of usable twist multiples is much affected by these considerations; a
typical set of twist multiple curves is given in Fig. 7.13. Generally the twists are
higher than for ring yarn and the combination of the higher twist and the more
disorganized yarn surface creates a rougher hand. At one time this was of major
concern, but fabric finishing techniques have improved and the market has adjusted
for the differences; the lower cost outweighs the tactile disadvantages. As mentioned
earlier, end-breakage rates are, amongst other things, a function of rotor diameter and
Yarn tenacity (mn/tex)
Polyester/cotton
150
Acrylic
100
Cotton
50
0
3
4
5
TM (machine value)
Fig. 7.13 Yarn tenacity curves for rotor spinning as a function of twist
196
Handbook of yarn production
speed. Whereas the larger rotors used in the 1980s gave minimum end-breakage
rates/lb at about 70 000 r/min, the newest small rotors (down to 28 mm are reported)
have minimum breakage rates at over 100 000 r/min. The minimum twist level achievable
is of interest because low twist yarns have a good hand; also, the spinning machines
have higher productivities at low twists. Generally, the minimum twist diminishes
with rotor speeds up to about 70 000 r/min and then levels off; under some circumstances
it rises at higher speeds. The lowest value of twist is a function of the radius at the
base of the rotor groove and the type of navel in use. The navels might be grooved or
non-grooved; they might be of steel or ceramic. Generally, the higher the rotor speed,
the less need there is for grooved navels. As previously mentioned, ceramic materials
are used to increase the life of the navels. The combination required for a given
product is often initially determined by a manufacturer’s recommendation, which is
followed by trials to find that best suited for the job.
7.2.8 Rotor bearing system
A rotor spinning machine has multiple rotors driven by a single tape and, because of
the very high speeds involved, special bearing arrangements are necessary. A common
design is to support each rotor on an assembly of pulleys, which rotate at speeds
lower than that of the rotor. Normal ball races are unable to survive at the highest
rotor speeds now used. Air bearings have been tried, with various degrees of success,
but the type of rotor support system most common now is similar to that sketched in
Fig. 7.14. The supporting rubber-tired pulleys are mounted on ball races and these
disks can safely rotate at the lower speeds.
Belt loading provided by jockey pulleys
(not shown)
Rotor shaft
F
Rotor
The bearings of the pulleys
with tires (not shown) provide
a reaction to the transverse
belt load (F).
V
Fig. 7.14 Rotor support system
Belt
Short-staple spinning
197
7.2.9 Winding on a rotor spinning machine
The winding function of a rotor spinning machine is separate from the rest. All that
is required is that the yarn be taken up at a constant rate. The rate of yarn removal is
determined by the surface speed of a pair of take-up rollers. The gearing between the
fluted sliver supply roller and the yarn take-up roller determines the machine draft
ratio. The yarn can be stretched involutarily, in which case there is a difference
between machine and actual counts. The speed ratio between the rotor and yarn takeup roller determines the twist level. Unlike a ring frame, the winding and twisting
functions are divorced and this permits the building of large yarn packages whose
size is limited only by the capability of the winder. As previously mentioned, the yarn
is then usually wound on a large cross-wound cheese that might (when fully built)
weigh some 10 lb (≈ 4.5 kg). Some rotor machines are capable of producing yarn cones
of about the same size. Use of these large packages reduces the yarn handling costs.
Rewinding (comparable to the winding from bobbin to cheese or cone in ring
spinning) is usually unnecessary in rotor spinning because the number of yarn faults
per package is usually low. However, it is difficult to anticipate which rotor will
develop a fault, and when a faulty condition arises a great deal of yarn can be made
before the fault is discovered. A cheese or cone running on the drive roller for long
periods without yarn being laid becomes damaged by the drive surface. Therefore,
some rotor yarn is occasionally rewound, but care is needed in rewinding because the
yarn can be overstrained in the process. A better alternative is to monitor the rotor to
detect a fault or an end-break immediately after occurrence and thus prevent the
building of a bad package. Commercial devices cause the package to become disengaged
with the drive roller when triggered by a fault; this is to prevent damage due to
abrasion of the unchanging surface of the package during a non-productive period.
Winding on an open-end spinning machine differs from winding on a separate
frame. In the former case, the yarn feed rate is set by the take-up roll at a constant
value, whereas in the latter case the yarn is supplied on demand. Compensation for
the change in length of yarn between the guide and the lay-on point on the package
is needed in open-end spinning. A simple scheme is to use a bow such as is described
in Section 9.1.6. The yarn diverted by the bow approximately compensates for the
change in length just mentioned and preserves a reasonably uniform yarn tension.
When a cone is being wound, a yarn storage system is required which is capable of
compensating for the differing wind-on speeds as the yarn traverses between the
small and large diameters of the cone.
The winder should control the yarn tension and maintain a uniform package density.
In addition, a pattern breaker is required to remove unwanted variations in package
structure, which arise as the diameter reaches certain critical measurements. Furthermore,
cradle pressure control is required to compensate for changes in package weight as
the package grows in size. A cradle lifter is often used to relieve pressure when an end
breaks (this prevents scuffing of the surface of the package). Sometimes the yarn user
requires a waxing attachment to apply paraffin (wax) to the yarn; this is particularly
true where the yarn is for a knitting application. Further discussion of winding is
given in Chapter 9.
7.2.10 Automation
Rotor machines are currently available that are capable of running at 130 000 r/min.
The maximum depends on the fiber being spun. Cotton can be spun at the highest
198
Handbook of yarn production
speed, acrylic at 20% less, and polyester and polyester/cotton blends at about 35%
less (90 000 r/min). It is impossible to use manual piecing at such high speeds and
automation becomes an operational necessity.
The rotor machine lends itself to automation, and patrolling robots that piece,
clean, and doff are commonplace. The robots follow a track round the machine. One
sort of robot patrols, opens up rotors, cleans them, and re-pieces according to program
or need. Another sort patrols and doffs when required. Frequently, machines have
automatic start-up programs and built-in monitoring systems that will read out the
machine performance over any reasonable period.
Automatic doffing requires that a supply of empty tubes have a starter yarn bunch
wound on each of them for start-up with the automatic piecer. Thus, there has to be
a supply of empty tubes, a bunch winder, and a transport and loading system, as well
as a system capable of removing the full bobbins in a safe and effective manner. The
latter has importance because a damaged cone or cheese is rarely recoverable and it
represents a considerable waste of effort and money. Damage is not restricted to a
violation of the surface but also includes the loss of effective transfer yarn tails,
which are so important to yarn users who tie packages nose to tail to reduce their
package handling costs. Each package should have a starter and finisher tail disposed
in a standard manner so that other machinery can find them.
7.2.11 Piecing in rotor spinning
Before an acceptable piecing can be made, the rotor has to be cleaned. Manual
cleaning involves stopping the rotor, opening the front cover, cleaning, shutting the
cover, and restarting; a time-consuming endeavor. Automatic methods of fulfilling
these functions are now standard to all new machines. The dirt in the rotor is usually
loosened by one or more blasts of compressed air through ducts in the rotor cover, or
the application of a scraper, and suction carries the debris away [13]. If a scraper is
used, the blade is made of a soft material to prevent damage to the rotor and the
sacrificial wear of the blade means that it has to be replaced periodically.
Automatic piecing requires careful control of the opened fiber entering the rotor
at start-up and the introduction of an end from the winder. After the inside of the rotor
has been cleaned, new fiber is introduced and a ring begins to build up. To explain
this, consider the series of pictures in Fig. 7.15. The end of the piecing yarn is shown
as square cut. The steps are exaggerated for the purpose of explanation. It is assumed
that the rotor has just been cleaned and that the thickness of the lines represents the
number of fibers in the cross-section. In diagram (a), a starter yarn (or piecing yarn)
is shown approaching the rotor groove; the fiber supply has just been started and a
thin ring of fibers has been laid in the rotor groove. The yarn is sucked in by the
vacuum and the air inside the rotor exists as a vortex. Thus, the yarn rotates about the
rotor axis at a lesser speed than the rotor and, because of the twisting actions already
described, the yarn end also rotates about its own axis. As the yarn is fed still further
into the rotor (diagram (b)) the end is laid in the rotor groove and it tries to rotate and
entangle the fibers already in the groove. The entangled end breaks into the fiber ring
(diagram (c)). At an appropriate time, the yarn is withdrawn from the rotor at an
appropriate speed. As the yarn is further withdrawn, the fibers continue to be supplied
to the rotor groove and extra layers of fiber are laid over the break (diagram (d)). This
process continues until equilibrium is reached. Meanwhile, the piecing contains a
thick spot followed by a thin one and this is not acceptable. This is why the end is
conditioned.
Short-staple spinning
199
Thickness of fiber ring at time t
Piecing yarn
Full groove
(b)
(a)
(c)
Groove fills during the
piecing operation
(d)
Fig. 7.15 Stages in piecing a rotor
For the purposes of explanation, let the circumference of the fiber ring be unzipped
to permit linear drawings (Fig. 7.16). The timing of the piecing is important, especially
at high rotor speeds. Thin spots break under the high tensions and, unless accurate
timing is used, it is impossible to get spinning going. Because a human operator
cannot reliably synchronize the stages in piecing at very high rotor speeds, automatic
systems become a necessity. If a square cut piecing yarn is introduced into the rotor
too soon and the rotor ring is thin, a very thin spot is generated in the piecing as
shown in Fig. 7.16(a). If the yarn is introduced late and the rotor groove is nearly full
to its normal operating level, a very thick spot is created. If the yarn withdrawal is
started too soon, there is a thin spot, and if it is started too late, there is a thick one.
Such a piecing is never perfect but it can be improved upon by tapering the end of the
yarn introduced into the rotor, because this reduces the sudden changes in linear
density, as shown in Fig. 7.16(b). Consequently, yarn end tapering is automatically
carried out by the machine, often using a pneumatic stripping device. At first sight,
some of the profiles shown do not seem to differ much. However, if the ordinates are
plotted to include all three components at any point, the result is quite surprising;
diagram (c) shows two such plots of the data for conditioned ends. The bad case
shown is for early yarn introduction, which resulted in a distinct thin spot; the other
case is a good piecing.
There is another good reason for conditioning the sliver end before piecing. Endbreaks occur randomly and the time between the break and the piecing varies. While
the end remains unrepaired, the combing roll is working the sliver end about to be fed
to the rotor. Fiber alignment, as well as linear density of the fringe, is affected. To
achieve standard conditions, the overworked sliver end should be discarded. The
easiest way to do this is to feed sufficient fibers into the rotor in the normal way and
then clean it before piecing. The machines are now designed to do this automatically.
200
Handbook of yarn production
Early yarn introduction
Fiber in groove at break-in
Piecing yarn
Added fiber
Late yarn introduction
Added fiber
Early yarn withdrawal
Added fiber
Late yarn withdrawal
Added fiber
(a) Square-cut yarn ends
Early yarn introduction
Fiber in groove at break-in
Added fiber
Late yarn introduction
A good piecing
Added fiber
Added fiber
Early yarn withdrawal
Added fiber
Late yarn withdrawal
Added fiber
(b) Conditioned yarn ends
n
A good piecing
n = linear density
Early yarn introduction
(c)
Fig. 7.16
Piecing diagrams
Some modern start-up devices reduce the machine speed for piecing. Other devices
assign the piecer to control the initial yarn withdrawal and synchronize it with the
accelerating machine before handing over control to the winder.
7.2.12 Fiber requirements
Bridging fibers can cause deterioration in yarn performance. The probability of a
bridging fiber is the ratio of the fiber length and the circumference of the rotor; thus
long fibers used in small rotors produce poor yarn. Conversely, within limits, short
fibers can produce reasonable yarn. (Comber noils can be spun successfully in rotor
spinners, which fact indicates that fiber length is not of paramount importance.)
Short-staple spinning
201
Some experimenters have stated that removal of noil from the feed material by
combing has little beneficial effect on yarn quality whilst others say that combing
improves yarn quality sufficiently to justify its use. There is an optimum fiber length
beyond which no further increase is beneficial; the precise value is determined by the
rotor size and geometry as well as the nature of the fiber. In making medium and
coarse yarns, some variability in length may be acceptable but it is worthwhile
controlling the short fiber content when spinning fine yarns. Deussen [13] points out
that the effect of fiber length should not be underestimated. Man-made fibers are cut
at 1.25 inches for coarse and medium counts whereas 1.5 inches is preferred for fine
counts. (For ring spinning, the standard fiber length for man-made fibers at all counts
is 1.5 inches.) Poor length uniformity can degrade the yarn quality but not so much
as with ring spinning. Highly crimped fibers tend to aggregate and behave as longer
ones; generally, high crimp is undesirable. Fiber finish also plays a part but, although
low friction finishes are desirable up to a point for man-made fibers, too slick a finish
can lead to yarn unevenness. Distinction must be made between fiber-to-metal friction
and fiber-to-fiber friction. The former should be at a minimum but the latter should
be high enough to prevent fibers slipping within the yarn structure when it is stressed.
Generally, rotor spinning is now confined to short-staple spinning with relatively
fine fibers. Fiber fineness plays a part not only in yarn strength, evenness, etc., but
it also affects the hand of the product. For this reason, fine cottons have become quite
popular with some rotor spinners. Experience has taught that a minimum of about
100 fibers is required in the yarn cross-section and the normal spin limit8 for cotton
varies between 24s and 50s cotton count (Ne). The spin limit is related to yarn
strength; higher strength fibers tend to reduce the end-breakage rate and increase the
spin limit (this is because it is the minimum strength of the strand that determines
whether or not a yarn breaks rather than the minimum linear density). End-breakage
rates can vary from 15 to 150 ends down per thousand rotor hours.
The practice in cotton marketing is to use micronaire as a measure of fineness, but
this clouds the issue of spin limit because micronaire is affected by fiber maturity as
well as fineness. The range of fiber finenesses now available for the man-made fiber
components makes closer matching possible with natural fibers to make good blends.
This is particularly true of polyester/cotton blends, which form a significant part of
the market. Man-made fiber development has led to the production of finer fibers
and the fineness has gone from the standard 1.5 denier to a range that includes values
below 1 denier. The fine fibers are prone to nep but they spin reasonably well in rotor
spinners except at the highest speeds. There is a tendency to break the finer fibers
during spinning, which was demonstrated in a study by Looney [22]. Reducing the
fineness of polyester fibers from 3 to 1.5 denier increased fiber breakage from 5% to
15%.
Fiber strength is an important parameter in rotor spun yarn because, in part, of the
comparisons made to ring yarns. A graph of the relationship between a particular set
of cottons and the yarns made from them is given in Fig. 7.17 as an example. (Some
people use the term ‘C × S’, which means yarn count multiplied by skein strength,
and it has the dimensions of tenacity.) At twist multiples higher or lower than the
optimum, the yarn strength declines. The cottons shown varied in length from 11/32 to
8 Spin limit is the count, or linear density, of the finest yarn which can be spun under the
prevailing conditions. There are other definitions.
Handbook of yarn production
Yarn skein strength (C × S)
202
Fig. 7.17
4000
NB C × S is known as the countstrength product and the units have
the dimensions of tenacity.
3000
2000
25
30
35
40
Cotton fiber tenacity (gf/tex)
Optimum rotor yarn tenacity as a function of fiber tenacity
13/8 inch (≈ 26 to 35 mm), the TM varied from 3.9 to 5, and the shortest fibers gave
the lowest yarn strength. The most important attributes for natural fibers used in rotor
spinning are fiber tenacity, fineness, and cleanliness, in that order. Fiber tenacity is
reflected in the yarn strength, and rotor yarns tend to be weak. Fine fibers work better
in rotor spinning than coarser ones. Dirty fibers create a problem in keeping the
rotors clean. Polyester, especially when an optical brightener such as titanium dioxide
is used, is somewhat aggressive and creates more wear than cotton. For man-made
fibers, cleanliness includes freedom from excessive fiber finish, debris, and oligomer.
With man-made fibers and blends of these with cotton, the factors can be ranked in
the order: fiber finish, tenacity, fineness, and length.
An area of interest in rotor spinning is the spinning of acrylic fibers, which, with
some machines, can be spun into yarn with a surprisingly low twist multiple. The
yarns produced are soft and of interest to knitters who want even, knotless yarns with
a soft hand and where strength is not of great importance. For this market, it is
possible to make the yarns economically because the low twist permits high production.
Economic success has been reported at counts as high as 30s cotton (45s worsted or
20 tex).
Another area of interest is the spinning of waste fibers. For example, cleaned
comber noils provide a cheap source of fiber and, in certain markets, the strength is
acceptable, but the dust must be removed before spinning because combing does not
remove dust from the noils. In some markets, noil is added to virgin fiber and
produces acceptable yarns; in some cases 100% noil can be used. The use of noil in
this way is of particular interest to makers of combed ring yarns because of the ready
availability of noil.
Blend yarns do not derive proportionate strength from the tenacities of the fiber
components because fiber elongation also plays a role. Each component may be
assumed to be extended by the same amount when a yarn is elongated. The load in
each fiber is a function of both the elongation and the modulus of the fiber. A yarn
made of a blend of stiff, weak fibers and extensible, strong fibers will usually fail
when the weak fibers reach their breaking elongation. At that point, a strong fiber
will only have contributed part of its strength to the composite at the time of failure.
Thus, for example, a polyester/cotton blend produces an effect such as is shown in
Fig. 7.18 [23]. However, if the percentage of weak fibers is small, the strong fibers
might be able to bear the entire load when all the weak fibers have failed. This is not
the usual case.
Short-staple spinning
Tenacity (g/tex)
30
203
Ring
20
OE
10
Polyester (%)
0
0
20
100
80
40
60
80
100
60
40
Cotton (%)
20
0
Fig. 7.18 Blend yarn tenacities
7.2.13 Maintenance
High speed precision machinery needs scheduled and thorough maintenance. It has
been suggested that 2 to 3% of the initial cost of the machine should be spent on
maintenance each year. Rotors and combing rolls are supposed to have operating
lives of up to five years but lack of proper maintenance, or abuse, can significantly
shorten useful life. The type of fiber being spun also has an influence. If the fiber is
cotton, then freedom from silica contamination is very important; if it is polyester,
the use of TiO2 or any other abrasive additives is a factor.
References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
Klein, W. A. Practical Guide to Opening and Carding, Manual of Textile Technology, 2,
Textile Institute, Manchester, UK, 1987.
Klein, W. A. Practical Guides to Combing & Drawing and Spinning, Manuals of Textile
Technology, 3 & 4, Textile Institute, Manchester, UK, 1987.
Schiffler, D. A. Roll Wraps in Ring Spinning: Part II, Effect of Fiber and Spinning Frame
Variables, Text Res J, 1993.
Shaw, J. Short-staple Ring-spinning, Text Prog, 12, 2, 1982.
Lünenschloss J. Textil Praxis, 22, 689, 760, 1967.
Stalder, H. Possibilities for Increasing the Productivity of the Ring Spinning Frame, Textile
Machinery: Investing for the Future, Textile Inst Ann Conf, 1982.
Herdtle, M. Fadenbrucherfassung, Minimierung und Behebung, Melliand Textilber, 6, 1984.
Barella, A. Yarn Hairiness, Text Prog, 13, 1, 1983.
Lünenschloss, J, Külter, H and Hoffman, B. Der Einfluss der Spindlexzentrizität auf das
Fadenbruchverhalten und die Garneigenschaften, Forschungsberichte des Landes NordrheinWestfalen, Nr 2417, Westdeutscher Verlag, 1974.
El Mogahzi, Y. Private communication, 1994.
Rohlena, V. Open-end Spinning, Elsevier Scientific Publishing Company, Oxford, 1975.
Hunter, L. The Production and Properties of Staple-fibre Yarns Made by Recent Developed
Techniques, Text Prog, 10, 1/2, 1978.
Deussen, H. Rotor Spinning Technology, Schlafhorst Inc, Charlotte, N.C., USA, 1993.
Siersch, E. Ein Beitrag zum Mechanismus der Fasertrennung und des Fasetransportes beim
OE-Rotorspinnen, Fortschritt Berichte der VDI Zeitschriften, 3, 56, 1980.
204
Handbook of yarn production
15.
Lawrence, C A and Chen, K Z. A Study of the Fibre-transfer-channel Design in Rotor
Spinning, J Text Inst, 79, 3, pp 367–408, 1988.
Lünenschloss, J, Coll-Tortosa, L and Phoa, T T. Die Untersuchen der Fäserströmung Im
Faseleitkanal einer OE-Rotorspinnmaschinen, 24, 355–8, 478–85, Chemiefasern TextileIndustrie, May June 1974.
Stalder, H. Faserauflösung und Faserführung beim Rotorspinnen, Vorträge anlässlich der 3.
gemeinsamen Tagung der Aachener Textilforschunginstitute, 1977.
Vila, F, Pey, A and Barella, A. A Contribution to the Study of the Hairiness of Cotton Openend Spun Yarns, Part 1, J Text Inst, 2, 55, 1982.
Vila, F, Pey, A and Barella, A. A Contribution to the Study of the Hairiness of Cotton Openend Spun Yarns, Part 2, J Text Inst, 73, 124, 1982.
Barella, A. Yarn Hairiness, Text Prog, 13, 1, 1983.
Landwehrkampf, H and Schreyer, F. USP 3 524 312, 1970.
Looney, F S. Mechanics of Open end Spinning, Private Communication, c 1975.
Lord, P R and Grady, P L. Proc 15th Canadian Textile Seminar, p 33, Kingston, Canada,
1976.
16.
17.
18.
19.
20.
21.
22.
23.
8
Long-staple spinning
8.1
Introduction: Effects of lengthening the staple
Long-staple spinning was originally designed to work with wool and other long
fibers, whereas short-staple spinning had been designed for cotton. Not only is wool
much longer than cotton, it is much more variable in length. Wool fibers normally
vary between 3 and 18 inches (≈ 76 to 457 mm) in length and between 8 and 60
microns in diameter. Therefore, the systems for wool must be able to cope with a
rather wide range of conditions. It is also interesting that wool is naturally crimped
at between 6 and 24 crimps/inch (≈ 0.2 and 1.0 crimps/mm); this is comparable to the
levels introduced artificially in cut or stretch-broken tow. This means that one would
expect more diversity in the machinery and processes used for long-staple opening
and carding than found with the corresponding short-staple processing discussed in
Chapter 5.
Man-made fibers are now used in blends but much of the appeal of wool lies in its
special character. Many consumers are willing to pay high prices for 100% wool
yarns. This applies in both apparel and carpet markets, and consequently there remains
a healthy market in fabrics made from pure wool. This is in addition to the market for
blends of wool with man-made fibers. We can divide the field into two without losing
too much of the total breadth. The two fields, which will be the subjects of major
discussion, are the worsted and woolen systems.
The worsted process as used for wool has similarities with the total short-staple
process involved in ring spinning as discussed in Chapter 5 except that fiber cleaning
has to be totally different because of the greasy nature of wool. The mechanical part
of the worsted process applied to man-made fibers, like its short-staple counterpart,
involves relatively little, if any, fiber cleaning. The worsted yarns produced by this
system are twisted and have a much higher strength than the woolen yarns just about
to be discussed. (Note: ‘woolen’ refers to the process and not the wool fiber.)
The woolen system is a short process designed to make relatively inexpensive
yarns. Stages of drawing, combing, and roving are dispensed with and the yarns are
spun directly from a card cylinder. The loss of the multiple doublings coming from
drawing and combing is countered by paying great attention to fiber blending in the
206
Handbook of yarn production
early stages of the process. Use of the woolen system has declined over recent years
but a description is included because the technology contains the roots of many
devices used in other processing.
8.2
Wool fibers and their preparation
8.2.1 Further effects of lengthening the staple
It is possible to improve fiber cohesion in yarn by increasing the fiber length; this
allows a lower twist multiple to be used. The lower twist, in turn, permits higher
productivity and yields a softer hand. Thus, there are a number of advantages to
working with long-staple fibers as compared to the short-staple ones we have already
discussed in Chapter 5, but there are limitations. Long fibers are difficult to manipulate
and the chance of fiber damage increases with fiber length. There are interactions
with the design of the machine; under certain circumstances, fibers that should flow
past, will adhere to the surface of a roller and be carried round that surface, as shown
in Fig. 8.1. If the fiber length is larger than the circumference of the roller, a roll lap
is likely to be produced. The fiber may be trapped on the roller surface as shown in
Fig. 8.1(b) instead of being carried away by the linear fiber flow. If this happens,
other fibers tend to be caught by the trapped fiber and a ribbon of fibers accumulates
around the surface of the roll. The ribbon is tightly packed and continuously builds
up, often causing damage to the rolls; these are called lap-ups. Such lap-ups usually
have to be removed by cutting them from the roll; this, too, can cause damage. The
ratio of fiber length and roll circumference is related to the tendency to lap. In cases
where fibers are incompletely separated, it is the tuft length that controls the situation.
Consequently, if the length of a tuft is longer than the circumference of the roll, lapups can occur. Fiber finish and crimp can also strongly affect performance in this
respect and these parameters must be closely controlled. Generally, when the roller
circumference is about twice the fiber length and normal fiber finish and crimp are
used, an acceptably low incidence of lapping is obtained. Thus, long-staple systems
usually have large diameter rollers in their drafting systems. In addition the ratch
settings have to match the fiber length. It follows that long-staple drafting systems
are larger in all dimensions than their short-staple counterparts. Long-staple yarns
are usually heavier than short-staple ones; this makes it desirable to use larger bobbins
to reduce doffing costs and thus the twisting portion of the machine is also larger. An
increase in size is nearly always associated with a decrease in speed. Therefore, it
x″
x′
Fibers adhere to rolls at x′
(a)
Fig. 8.1
Fibers trapped under new layer of fiber at x″
(b)
Roller lapping
Long-staple spinning
207
should be no surprise to find that long-staple machinery is to a larger geometric
scale, and is slower, than short-staple machinery.
Long fibers tangle more easily than short ones (a man with a crew cut has much
less difficulty in combing his hair than does a person with long hair). Proper lubrication
of the fiber eases the problem. Crimped fibers tangle more easily than smooth ones.
The carding process aims to disentangle the fibers and separate them, but it is found
that the flat-top card damages long fibers in attempting to disentangle them. Consider
the carding action between two sets of teeth moving at different speeds in about the
same plane. One fiber end is caught in one set of teeth and the other end is caught by
the second set. Figure 8.2(a) shows a typical fiber snaking between the teeth on a
surface from A at one end to B at the other. If there is already a tension acting at A,
then the tension at B is given approximately by applying Amonton’s Law.1 Appropriate
angles as well as the input and output tensions are as shown in diagram (b). It will be
seen that the more sinuous the shape of a given fiber, and the closer the pins are set,
the greater will be the accumulation of tension. In practice, the tension at A is
generated by interfiber friction in the fiber supply, or an interaction between fibers
and pins in a preceding section, or both. That at B is X times as much. A factor Z =
Tout/Tin is due to the fiber reactions against the pins just described, but another factor
Y is affected by the fiber crimp and stiffness. (Hence the approximately equal symbol
in the equation in the footnote.) The precise relationships between X, Y, and Z are not
known. The reactions are caused by the in-built tendency for the fibers to take up a
zigzag shape. Pressure from the teeth de-crimps the fibers as they are forced into the
interstices, and this produces a friction force that contributes to the tension at A. In
φ2
φ1
Motion of the fiber relative to the pins
T1
T2
B
T2
A
φ3
T1
φ4
(b)
(a)
Fiber
Feed roll
T
C
D
(c)
(d)
Fig. 8.2
Fiber tensions created by combing
1 Tout ≈ Tin eµθ where the angle θ = φ1 + φ2 + φ3 + φ4 and µ = coefficient of friction.
208
Handbook of yarn production
any event, the output tension at B is heavily affected by the input value at A. The
angle of wrap, the fiber crimp level, the pitch of the teeth, and the coefficient of
friction also affect it. The lubrication and crimp are of increasing importance as the
staple length increases beyond 4 inches. The cumulative angle of wrap is affected by
how many zigs and zags there are along the length of the fiber. Beyond a certain
critical length, fiber breakage will occur. Using a pinned feed as shown in diagrams
(c) and (d) can double the critical fiber length; such systems are commonly found in
long-staple machines.
8.2.2 Wool fiber cleaning in worsted mills
Wool is usually supplied to the mills as shorn or pulled fiber. Shorn wool is taken
from living animals and pulled wool is from carcasses. These untreated materials are
known as greasy wool because of the grease, suint (from old French ‘suer’, the verb
‘to sweat’), and other animal excretions that coat the fibers. Together with vegetable
and mineral particles, there can be up to 70% foreign matter in greasy wool. The wool
grease (or so-called ‘brown grease’) coating the fiber contains a waxy material called
lanolin, which is a valuable by-product useful in making ointments and cosmetics.
Suint is dried perspiration that contains valuable potassium salts. Greasy wool will
not yield to purely mechanical means of cleaning and therefore it is necessary to
scour the fibers before further mechanical treatment. There were relatively few
developments in the first half of the twentieth century; however, the problem then
started to receive more attention, especially in the grower countries.
The traditional method of scouring is to divide the greasy wool and feed it to a fork
frame (Fig. 8.3(a)). A procession of forks carries fiber through a succession of liquor
troughs (or bowls) in which the fibers are washed and rinsed. The frame may be some
60 ft (≈ 18 m) long. The scouring liquor is traditionally an alkali (soda) soap solution,
since it readily permits the recovery of lanolin, but this requires soft water, otherwise
there are troublesome insoluble lime deposits. There is a diversity of ways of running
the liquor through the frame and we can only describe one. Care has to be taken to
control the temperature and the pH level of the liquor, otherwise the fibers will be
damaged and the dyeability of the fibers affected. The degree of alkalinity or acidity
is recorded by pH readings. Generally, the temperatures and pHs of the first stages
can be kept high because the grease tends to protect the fiber. As the protective grease
is removed from the fiber in its passage downstream, the temperature in successive
bowls is reduced and the pH level is moved nearer neutrality. The final bath is merely
a rinse in plain water, after which the wool is passed through squeeze rollers to
remove most of the moisture. An average time for the wool to pass through the fork
frame is 8 minutes. Scouring tends to make the fibers brittle unless they are oiled and
the fibers are kept below about 120°F (≈ 50°C) during scouring.
The wool grease has to be separated from the scouring liquor to allow a reasonable
run of the equipment. Once again there is a diversity of systems and it is only possible
to describe an example in this book (see Fig. 8.3(b) for a simplified diagram). The
liquor still degrades, and the scouring liquor has to be completely changed every few
days. In addition, there is a need for liquor make-up and fresh water to rinse, and thus
the process is a great consumer of water. Solids, such as sand and other heavy
particles, are allowed to settle in tanks in the scouring liquor circuit and these solids
may be intermittently pumped out or otherwise removed. Fiber loss into the scouring
liquor causes a problem because fibers tend to bind the contaminants, and when they
Long-staple spinning
209
Greasy wool input
Direction of harrow forks
Scouring
liquor out
Scouring
bowl
Scouring
liquor in
Filter
Settling tanks
Filter
Centrifuge
Squeeze-roll
system
Rinse
water out
Rinse bowl Rinse water in
Replenishment
Hot air + moisture out
liquor
(a) Fork frame
Cool air in
Recirculated and
Heavy sludge
cleansed liquor
removal
Waste
Clarify
wool
grease
Crude
lanolin
Fiber
transfer
Scoured
wool
output
Heaters (not shown)
(c) Dryer
(b) Scouring liquor treatment
Sketch not in proportion, so as to illustrate the principles more clearly.
The drawings have been simplified. For example, only one scouring bowl is shown.
Fig. 8.3
Fork frame
decay, they become malodorous. Fatty substances contaminated with fiber and other
light substances tend to float to the surface in the settling tanks, and sludge centrifuges
may be used to remove these lighter solids continuously. (Centrifuges are machines
for separating solids from liquid suspensions.) Disposal of the liquor not processed
by the recirculation system is discussed in Section 8.2.5. The wool grease is separated
from the contaminated fatty solids and clarified (see Fig. 8.3(b)), to be later refined
by the purchaser. If a market exists for potash at a price sufficient to cover the
costs of separation, the salts are converted to whatever intermediate the buyer will
accept.
Some attention was given to solvent scouring but no significant market penetration
was achieved. The use of detergents was once claimed to give whiter, loftier, and
stronger fiber; detractors’ claims and counter-claims were not so significant as the
impact on costs and the convenience of being able to scour to the desired pH level.
In recent times, the development of effective detergents caused costs to become
competitive with alkali solutions. Whichever method is used for scouring, the main
aim is to remove the wool grease at minimum cost with as little damage to the fiber
as possible. Cost considerations are a large factor in the choice of method.
An important past problem was the felting of the wool during scouring, but suitable
chemical treatments are now available to reduce the ensuing shrinkage [1]. Felting
occurs because scales on the wool fiber act as ratchets, which favor movement of the
210
Handbook of yarn production
fiber in one direction rather than the other. The result is that wool structures tend to
become densely packed as the mass shrinks due to the relative movement of the
fibers. Felting makes the disentangling of fibers difficult.
Some machines were developed in the mid-twentieth century to improve performance
by increasing productivity and reducing the tendency to felt; one of these is the
suction-scour machine. In this machine, wool passes round the lower surface of a
number of horizontal, perforated drums rotating in the scouring bowl. A stream of
scouring liquor flows from the outside to the inside of the drums and the wool floats
near the surface of the liquor in a stream flowing from one drum to the next.
The wool leaving the scouring plant should have a moisture content of less than
40%. This means that, where aqueous scouring systems are used, the wool has to be
passed through at least one pair of squeeze rollers (mangles) to reduce the excess
moisture. The damp wool is then passed through a dryer in which hot air is circulated
through the blanket-like layer of wool carried on a lattice conveyer. Cool air is
admitted at the delivery end and is progressively heated as it passes through the fiber
as shown in Fig. 8.3(c). This enables the dried wool to be delivered in a cool state.
After chemical treatment, it is essential to oil the fibers before proceeding with
further mechanical treatment. Scoured fiber is usually hot air dried and oiled to a
level of 0.75%2 to facilitate further working. It is interesting to note the need to oil
the fiber after the surface has been rather drastically cleaned. Use of the wrong oil or
the wrong quantity of oil can result in processing difficulties. See Section 8.2.4 for
further discussion.
Where the vegetable matter among the scoured and oiled fibers is present only in
moderate quantity it can be dealt with at the card. Dirtier fibers need some precleaning. A partial alternative to the mechanical removal of vegetable matter is to
destroy it by chemical action. Sulfuric, hydrochloric, or other acids (such as those
produced by salts such as aluminum chloride when heated) may be used. These
treatments reduce the unwanted matter to carbon and the process is known as carbonizing.
Usually, the dirty scoured wool is steeped in acid, dried, and baked. Drying is
accomplished by draining, mangling, and heating. The particles are then crushed and
beaten out before diluting the acid and then neutralizing by soda washes. The dilution
is necessary to prevent overheating the fibers due to the exothermic reaction during
neutralization. Rinsing and drying follows it. Such carbonization gives some loss of
fiber amounting to 5 or 6%. Also, there can be some loss in fiber strength, particularly
if high drying temperatures are used. At 250°F (≈ 120°C), the loss in strength can be
as much as 30%, but at 100°F (≈ 38°C) the loss is negligible. Again, it must be
pointed out that there is a necessity to oil the fibers after chemical treatment to
prevent fiber breakage and electrification.
It will be seen that the wet cleaning process is complex and it can be added that it
is expensive to both install and run. This alone tends to make wool products expensive;
consequently a great deal of care is usually taken to preserve quality.
The first mechanical cleaner in line is called a burr picker. A second cleaning is
achieved by blowing air through the burr picker to remove light dust and impurities.
One of the last opportunities mechanically to remove all but the last remnants of
vegetable matter is in the carding operation. These mechanical operations are described
in Section 8.3.2. Any residual vegetable matter left in worsted slivers is usually
2 A higher level might be used occasionally, especially if obsolete Noble combers are in use.
Long-staple spinning
211
removed in the combing process. An extensive discussion of the variety of processes
is given in the Wool Handbook [2].
8.2.3 Fiber cleaning for woolen mills
Many traditional woolen mills use recycled fibers and then there is no need to use
methods described in the previous section, but there is a need to clean and sterilize
the fibers. A shortened cleaning system is then used. However, if the mill wants to use
virgin wool fibers, then a cleaning system such as that just described is required for
at least part of the intake of fibers.
8.2.4 Fiber lubrication
The importance of proper fiber lubrication has already been stressed but it should be
realized that considerable care has to be taken in the choice of oil and the method of
application. There are many oils used (and even more claims, many of which are
unsubstantiated) and such materials as olive oil plasticize, lubricate, and reduce
electrification. Over the years, many competing oiling formulations have been derived
and it is beyond the scope of this book to discuss them further. The oil is usually
applied to a layer of fiber, another layer of fiber is added, then more oil is applied, and
so on. Too much oil, or one that is too viscous, or one that becomes sticky over time,
causes lap-ups and chokes with the result that the fiber becomes unworkable. Also
insufficient lubrication or lack of plasticizer can cause an undue amount of fiber
breakage in the mechanical operations. Such fiber breakage impairs operational
efficiency and reduces the quality of the product.
Not only does the oil change the coefficient of friction, but it also reduces the
tendency for the fiber to charge electrically. Insulated surfaces sliding over one
another create electrical charges. All fibers are subjected to sliding contact with other
surfaces during processing and thus the ‘oil’ has to possess some ability to allow an
electrical current to flow to minimize the charge. If not, since like charges attract
each other and unlike ones repel, the results are that (a) the fibers cling together and
they are difficult to separate and (b) some fibers adhere to machine parts and cause
uncontrolled fiber flow and breakage. Such fiber behavior can make processing
exceedingly difficult and increase the probability of lapping and tangling. (It might
be remarked that these situations provided the early makers of man-made fibers with
some valuable insights as to what they had to do in the way of fiber finish to make
their fibers capable of being worked.)
The presence of a sufficiency of proper oil can reduce the need for a high atmospheric
humidity. Humidity also reduces the tendency towards electrification of fibers. For a
woolen system with only small amounts of oil, the rh has to be at least 65% and a
humidifying system becomes a necessity in many climates. Air at 65% rh and 70°F
(≈ 21°C) holds about 5.2 grains/cu ft (≈ 12 g/m3). The quantity of water that has to
be sprayed into the atmosphere can be quite large and the cost significant. Table 8.1
gives some typical rh values at various processing stages in the worsted system. It
will be noted that the rh values needed for carding and spinning in the worsted system
are considerably lower than the 65% just quoted. Wool fibers in the worsted system
are oiled.
When different sorts of fiber are to be blended, it is desirable that each sort of fiber
be lubricated separately. Many man-made fibers already have fiber finish applied by
212
Handbook of yarn production
Table 8.1
Typical % rh in processing of wool
Mixing and blending
Carding
Combing
Drawing (French)
Spinning
% Min
% Max
65
60
65
65
50
70
70
75
75
55
the fiber maker, whereas wool is scoured before mechanical processing. Most manmade fibers process reasonably well at 55% rh and thus the further importance of
oiling the wool can be seen. Use of appropriate moisture contents and oiling help to
keep the wool fibers pliable and resistant to rubbing. It has been explained that the
lubricant has to perform some important functions but there are other requirements.
The oil must be capable of being easily removed after it has fulfilled its purpose; it
should not degrade, stain, or otherwise mark the fiber; it should not damage the
machine in any way. It should be compatible with the finish on any man-made fiber
used with it; both should be chemically stable under a variety of conditions during
prolonged storage; and it should not produce any fire or health hazards.
8.2.5 Disposal of wastes from scouring
Disposal of the wastes from scouring is still a problem and traditional methods are no
longer acceptable. Waste water from scouring is a particularly difficult effluent to
deal with because it contains both organic and inorganic components. Inert sediments
and vegetable matter produce little problem in this respect, but the wool grease and
soluble organic salts do have highly significant effects. The polluting effect of such
aqueous wastes is indicated by its biochemical oxygen demand (BOD) to achieve the
decomposition of organic waste by aerobic bacterial action. The rate of discharge of
these waste materials can thus be measured in pounds of BOD per day (or by the
equivalent of that generated by a number of inhabitants of a community). The BOD
rate depends upon the waste contents of the fibers, the level of production, and the
efficiency of the waste treatment systems used. Several methods of treating wastes
exist. Biological methods are low in cost and are capable of dealing with both soluble
and insoluble impurities. However, they are sensitive to variations in ambient
temperature, they can be adversely affected by poisoning, and they produce large
volumes of sludge. Gravity settling of the sludge in lagoons is used to produce the
biological action. If the sludge is spread on the land as a way of disposal, a large
acreage is required. Flocculation of the liquors has not proved to be satisfactory on
the small scale that most mills require. Evaporation of the water from the liquor
followed by condensation to recover the water is another concept that has been tested
and used at various places because it has the attraction of a reduction in water usage
and a high yield of valuable wool grease. The cost of water is a severe burden in some
areas of the world. An acid-cracking method is possible in which the suint liquors are
evaporated and the residues of salt are calcined to produce saleable potash. However,
the evaporation process can be expensive.
Long-staple spinning
8.3
213
Worsted systems
8.3.1 Long-staple processing within the worsted family of systems
Within the worsted family of processing systems, there are many variations. These
systems range from the traditional processes used for making high quality yarns from
wool, to short processes that use man-made fibers solely and require no carding or
combing. These systems use some of the same equipment at certain stages of their
manufacture and it is convenient to use the adjective ‘long-staple’ when the process
stage is useful over a wider range than might be inferred by the adjective ‘worsted’.
8.3.2 Worsted system carding
Worsted yarns are made from long, lustrous varieties of wool and they are usually
combed to improve the luster, smoothness, and strength of the yarn. Wool fleeces are
sorted into pieces of reasonably uniform quality; these sorts of greasy wool are
blended and, if necessary, deburred before being scoured. The scoured, dried wool is
conveyed by an airstream (or by a lattice conveyor) to a temporary storage bin, or
area, from which a lattice card feed draws fiber at a controlled rate. An automatic
weighing system is usually used to control the thickness of the fiber fleece fed to a
roller-top card (Fig. 8.4). Breast works, diagram (a), are used to feed fibers to the
first main cylinder of a card of the sort shown to small scale in diagram (c). It was
once thought that at least five licker-ins were needed to card wool, but the introduction
of ‘metallic’ wire has reduced the necessity for so many. The so-called metallic wire
refers to saw-toothed wire of the type illustrated in Fig. 5.12. Special metallic wire
(Fig. 8.4(b)) is used to keep burrs and other extraneous matter on the surface of the
cylinders and yet allow the fibers to be pressed between the rows of teeth. This
arrangement permits the removal of the unwanted vegetable material by bladed burr
beaters. Special wire-covered Morel cylinders work with burr beaters and are often
substituted for the redundant licker-ins, to provide extra deburring stages. Morel
beaters are equipped with special flat-topped wire to keep the fibers on the surface
and so enhance the cleaning function. The main sections of these cards give little
other opportunity for cleaning.
Burr beaters are in the range of 5 to 6 inches (say, 130–150 mm) diameter and run
up to 2000 r/min. Atkinson and Saunders [3] showed that eight-bladed burr beaters,
run at about 1000 r/min, were nearly as effective at removing seed/shrive and vegetable
matter as similar beaters running at 2000 r/min. Eight-bladed beaters were preferred
to two-bladed ones. Various blade profiles were used and high relief angles were
used. Burrs were removed at up to 80% efficiency and seed/shrive at up to 60%.
Morel clearance and surface fiber density are of importance. The wool in the waste
varies between 10% and 20%, the bulk being mostly vegetable matter. Efficiency of
removal of unwanted matter depends largely on the speed of the beater, up to a
limiting value of about 1800 r/min. Clearance between the burr beater and the Morel
has an effect on efficiency; opening up the clearance by 2.5 mm drops the efficiency
by about 15% [4]. On single swift cards, two Morel/burr beater pairs are placed in
tandem between the forepart, or breast works, and the swift [5]. Deburred and roughly
opened fiber is transferred from the last licker-in or Morel beater to the first main
cylinder (or ‘swift’).
The main cylinder carries the fiber to the first worker/stripper fiber combination
(Figures 8.4(b) and 8.5(c)). The surface speed of the worker (which is the larger of
214
Handbook of yarn production
Scoured
fiber input
Output fiber web on
first main cylinder
Divider
Divider
Stripper
Brush
Enlargement
Burr beater with
saw-tooth wire
Takers-in
Divider
Tightener
Worker/stripper sets
(b)
Divider
Tightener
Burr beaters shown with black fill
(a)
Swift
Doffer
Breast cylinder
(c)
Fig. 8.4
Roller-top card
the two rollers) is less than that of the cylinder, and the teeth are angled to pick up
fibers rather like a doffer. As the fibers picked up by the worker move away from the
cylinder, the rapidly moving cylinder teeth comb out the trailing ends of the retreating
fibers. The stripper runs at a higher surface velocity than the worker does, and this
causes fibers to be stripped from the worker and to be drawn and combed again
before being returned to the cylinder. Fiber can take various paths as indicated in Fig.
8.5(b). The surface speed of the stripper is less than that of the swift, so it acts rather
like a licker-in, and the teeth are angled to facilitate the fiber transfer. Typical relative
velocities of the various components, taken in the order that most fibers meet them,
are shown in Fig. 8.5(c). Within the normal range of settings between the wire
surfaces, there is surprisingly little effect on the production of noil obtained in later
combing [5]. However, the speed of the set is important in determining the noil
percentage which implies that it has an effect on fiber breakage. Robinson [6] suggests
that fine worsted-style wools can be carded at higher rates than normal, provided an
antistatic lubricant is used. It will be noted that worker/stripper sets replace the flats
used in short-staple processing. Worker/strippers not only fulfill a similar function of
dividing tufts, but also give greater longitudinal blending than can be achieved with
flats.
Long-staple spinning
215
6
7
Stripper
Fancy
Worker
5
2
3
To doffer
4
Cylinder
1
Cylinder
Wire size exaggerated for clarity
Alternative fiber paths:
1, 4, 8
1, 2, 3, 4, 8
1, 4, 5, 6, 7, 3, 1, 4, 8 etc.
(b)
The fancy roll raises fiber to the surface of
the wire to facilitate doffing
(a)
Velocity, log scale (ft/min)
104
Cylinder
8
Fancy
103
Strippers
102
Licker-in
Doffer
10
1
Workers
Feed
Stage
(c)
Fig. 8.5
Carding elements
The layers of fibers on the swift may be considered to consist of a component from
the feed and a component of recycled fibers, due to the inefficiency of the doffers.
The swift-to-doffer setting, and the speed ratio between them, affects the efficiency
of fiber transfer to the doffer. The fact that there is a choice of fiber path at each
worker/stripper means that there are longitudinal displacements of successive proportions
of fiber and this gives a useful blending action. Errors tend to be smoothed out by
these random lengthwise relocations of the fiber population. There are several stages
of working and stripping on each main cylinder. Normally, several main cylinders are
used in series in a normal carding set; consequently, there can be up to about a dozen
worker/stripper stages in the total process. Incomplete fiber separation and entanglement
of fibers in the material delivered by the card affects fiber breakage in later processes.
Web is transferred from one main cylinder to the next by means of an intermediate
doffer cylinder. Often crush rolls are used at this point, to reduce the size of any small
burrs left in the stock after the deburring phase of the process. Many of the crushed
burrs fall out from the fiber because the crushing tends to break up the spikes and
sharp edges of the burrs. Also, a Morel burr cylinder is often fitted at the transfer
point. Fibers leaving the last worker/stripper tend to be deeply embedded in the
cylinder wire, and it is necessary to raise the fiber to the surface to facilitate doffing.
A ‘fancy roll’ (Fig. 8.5(a)) carries out this function and the member has long, flexible
wire teeth. The surface speed of the fancy is higher than that of the cylinder, with the
result that the fibers are raised and brushed forward to be better caught by the doffer.
The doffer is of conventional design and operates much as was described earlier,
except for the condensing stage.
216
Handbook of yarn production
8.3.3 Man-made fibers in worsted spinning
The production of man-made staple fibers has been addressed in Chapter 2 but some
mention should be made of their use in worsted spinning. If the fiber is to be carded,
the man-made fiber is supplied in bales. The unopened bales of fiber should be
conditioned for two or three days before cutting the bands and removing the bale
wrappings. An rh of between 55% and 70% is usually needed for this purpose. Since
the fibers are clean, it is not necessary to risk nep creation by over-carding and ICI
recommended low carding rates [7]. The sliver should not be ‘backwashed’, otherwise
the risk of nep production in later processes is increased. Backwashing is a washing
procedure which will remove fiber finish from the man-made fibers. Changes in
frictional characteristics as compared to wool might require alterations in the back
draft in combing. It is considered inadvisable to add oil or other lubricant to the sliver
after combing. However, if the sliver is dyed or printed to produce special effects, the
fiber finish is removed and a suitable dressing becomes needed to control static
generation in subsequent processes. If the sliver becomes matted or compacted, a
preliminary opening in a gill box is suggested.
Where the man-made fiber is to be blended with wool, it is advisable to do so in
the preparatory gilling stages (Section 8.3.4) before re-combing. This implies that the
slivers of wool and man-made fiber have been produced separately. When intimate
mixtures of man-made fibers and wool are combed, mixed noil is produced, of variable
fiber proportions, and this is of lower value than unmixed noil. Color matching might
become a problem, but comb mixing is recommended for the production of high
quality tops (i.e. sliver). Differences in coefficient of fiber friction also affect the
slubbing (or roving) twist needed; lower values than those normally used for wool
might be employed. This has economic advantages. In spinning, different ratch settings
might be required from those used with wool. The spinning and folding twist levels
required for a given end use are affected by the fibers used. Differing levels are
necessary for the various man-made fibers, wool, and blends. Strength is not always
the main concern; hand and appearance are often more important.
8.3.4 Long-staple drawing
A traditional drawing operation (‘gilling’) is carried out on pin drafters. Faller bars
with pins in a comb-like configuration move similarly to the simplified fashion
shown in Fig. 8.6(a). Two sets of intersecting faller bars are usual. One set enters from
the top of the fiber stream, and the other enters from the bottom. There up to 100 such
faller bars in a machine and about 30 of them engage the sliver at one time. Studies
of pinning density and fiber loading have shown that increased pinning densities can
lead to irregularity, probably because of the increased drafting forces produced.
However, if the pin density is not increased to extreme levels, the pins act to control
unstable fiber movement in the drafting zones. At each drawing stage, a number of
slivers are creeled to form the input and this produces a useful doubling that improves
long-term regularity and blend. There are usually four stages of drawing before
spinning [4]: three stages of intersecting gills and a final stage on a rubbing finisher
or speed frame (i.e. roving frame). The gill frames have a restricted productivity of
about 2000 faller drops/min and there is an incentive to replace them with rotary
gills, or caterpillar drafting or the like. These newer systems obviate the noisy, rectilinear
motion of the faller bars. Also, caterpillar and chain gills are capable of a greater
length of contact in the draft zones as shown in Fig. 8.6(b). They operate at up to four
Long-staple spinning
217
Path of faller bars
Caterpillar
Sliver
Faller bars
Brush
Delivery rolls
Feed rolls
Back rolls
Brush
Front rolls
Caterpillar
(b)
(a)
Fig. 8.6
Gilling
times the productivity rates of the traditional gilling frames. Automatic doffing
(autodoffing) of the sliver cans is becoming well established with can sizes up to
60 inches (≈ 1.5 m) diameter.
The front roller system in a classical machine is a three-roll set. The drafting, feed,
and sensor rolls are often pneumatically weighted and sometimes pneumatic sliver
transport systems are used to carry the sliver to the front roll assembly. Sliver mass
sensor systems are fitted to new machines and the electrical signals are used as
computer inputs; the algorithms in the computer control the short- and long-term
linear density variations. Stepless drive motors allow tension controls, as well as the
differential speeds necessary to control drafts. Autoleveling systems are frequently
used, especially on the first stage of drawing.
8.3.5 Top making
It is more common to make long-staple sliver ready for spinning or drawing elsewhere
than it is in short-staple spinning. The combed wool sliver is a high value product and
it is worth the extra cost of transport if the material produced by a specialist is of
superior quality. Another reason is that the long-staple sliver is more durable because
of the use of long fibers. It is common to add as much value as possible before
shipping and the material shipped is known as a ‘top’. These tops are frequently
stored as a ball, in which the sliver is cross-wound onto an external package as shown
in Fig. 8.7. This is made possible by the high fiber cohesion in these long-staple
slivers.
8.3.6 Long-staple combing
Long-staple slivers are often combed to improve fiber orientation and the appearance
of the final product. Blends of natural and man-made fibers may be made by combining
ends of sliver in the drawing and combing processes.
Two main combing systems exist but the major one uses the Heilman or French
comb, which has a rectilinear, intermittent action somewhat analogous to the cotton
comb. The machines are larger than their short-staple counterparts and operate at up
to 240 nips/min. Alternatively, tops may be made from tow by stretch-breaking as
was described in Chapter 2. At the machine level, new arrangements of drive cams
218
Handbook of yarn production
Fig. 8.7
Sliver ball
combined with the use of lightweight materials have permitted increases in speed
without substantially increasing the noise level or adversely affecting the performance
of the machines. In modern machines, draw-off aprons control the emerging web,
draw-off cylinders are equipped with lap detectors, combs are fitted with cleaners,
and computers are used to control automation.
8.3.7 Long-staple roving and spinning
Roving and yarn may be made on equipment that is fundamentally similar to that
described for making short-staple yarns. Draft ratios are higher as is the fiber cohesion.
Fiber oiling has to be considered. As mentioned earlier, the increased fiber length in
worsted systems means that the size of the rolls and the distance between them are
correspondingly larger. Yarn counts are heavier and the yarn packages have to be
correspondingly larger to prevent undue loss in production due to doffing. This means
that larger flyer and ring sizes and lower speeds are used than in short-staple spinning.
In drafting, variable fiber length is often quoted as the most important factor in
producing irregularity; however, increasing fiber crimp has also been shown to have
a detrimental effect on the ends-down rate in spinning. There is an inference that high
crimp causes irregularities. High drafting forces lead to the need for higher roll nip
pressures in the drafting system, otherwise slippage between fiber and roll leads to
irregularity. High pressures between rolls and soft surfaces in a drafting system tend
to damage wool. Apart from the dangers of fiber lapping, it is desirable to use as large
a diameter drafting roll as possible because small rolls produce high pressures. The
diameter is limited, of course, by the roll setting and that, in turn, is related to the
fiber length. Historically, various twisting systems were used that included ring,
mule, and cap spinning; but the use of the latter two has declined and we need only
concentrate on the ring frame. It is interesting to note that better results are obtained
with increasing mean fiber length, but that variability in length seems to have little
effect. Fiber diameter and percentage of short fibers correlate with the frequency of
thick spots in the yarn [8].
8.3.8 Worsted spinning
In principle, long-staple ring spinning is similar to that described for short staple.
Consequently, the descriptions will not be repeated. Yarn tension in spinning is a
Long-staple spinning
219
function of the yarn count, spindle speed, ring size, and some other factors. With the
heavier counts in long-staple spinning, it is desirable to increase the ring size with the
heavier yarns, not only to reduce doffing costs but also to reduce the incidence of
knots or splices. To maintain an acceptable spinning tension, it is necessary to reduce
the spindle speed and/or reduce the balloon size. It is here that collapsed balloon
spinning finds its place (see Appendix 9). Differences in count and running conditions
lead to means of controlling fiber flow that differ from those already described.
Because of the wide range of products, there is a wider band of technology in longstaple spinning than in short staple. As examples: rings are sometimes lubricated;
traveler designs are more varied; and travelers can be made of polymeric, composite,
and steel materials. The large packages typical of worsted spinning require that the
bobbin length be large. This implies a significant variation in spinning tension during
spinning unless the bobbin is moved up and down during the wind, to keep the
balloon length about the same. Some machines do this and some oscillate both the
bobbin and ring rail for this purpose. The mechanical complication of this is repaid
by a reduction of the variation in tension, which, in turn, leads to a more consistent
product and a reduction in end-breakages.
Automatic winding has become established as has the use of electronic clearing,
but the joining of end-breaks and cuts during winding is rather more difficult than
with short-staple yarns. For medium length staple, splicing is feasible, as explained
in Chapter 9. Splicing can give an almost faultless join but knotting is still often used
for the heavier yarns. Many worsted yarns are plied and several forms of twisting are
used. Traditionally, worsted warp yarns have had to be plied to withstand the rigors
of weaving. Plied and quasi-plied yarns are made by ring uptwisting, two-for-one
twisting, Sirospun processes (see Chapter 10), novelty twisting systems, and others;
the products include not only simple plied but also certain fancy yarns. These fancy
yarns have spiral, loop, nub, bouclé, ratine, flake, chenille, or other special effects,
but they are too various and complicated to be explained here. In such a wide range
of yarns, the requirements for fault clearing in winding vary enormously. It is obvious
that the market expectations of a particular yarn must determine the levels at which
faults should be removed, and any system must permit variation according to need.
8.3.9 Semi-worsted and related systems
There are two main divisions in long-staple processing of natural fibers and these are
known as the worsted and woolen systems, both of which are based on systems
originally designed for wool. As new man-made fibers have been developed, the
tendency has been to blend them with wool, or to displace wool altogether (wool has
become a relatively expensive fiber). The wholly man-made, long-staple yarns have
become popular for carpets.
Attempts have been made to shorten process lines to gain some economic advantage.
This philosophy has been applied to the worsted system, and although the solutions
are somewhat diverse, it is possible to categorize them under the heading of ‘semiworsted systems’. The most popular use is in carpet yarn manufacture and the count
range is between 3s and 12s worsted. A typical system uses a one- or two-cylinder
card set with appropriate deburring stages, and there may be six or seven worker/
stripper combinations per card. Sometimes two doffers are arranged to give a split
web and the left and right portions are converted to sliver. It is possible to use
multiple doffers, each producing split webs. Care has to be taken when producing
220
Handbook of yarn production
multiple slivers to balance the outputs to keep uniform sliver weights. Such multiple
doffer split web cards are made at up to 100 inch width with productivities of up to
450 lb/hr. Often, two or three passages of drawing are used, according to the yarn
count (above 10s worsted, say, three passages would be used).
In the post-carding processes, the first drawframe normally has an autoleveling
attachment to even out the errors from carding. The first frame can have up to 12
slivers per head in the creel, so there is a considerable amount of doubling to improve
the long-term regularity. The intermediate and finisher drawframes might take three
slivers per head with a draft of about 8. Therefore, there is a progressive reduction
in sliver weight from about 300 grains/yd at the card to about 50 grains/yd at
the finisher drawframe delivery. The productivity of the first two frames is about
200 lb/hr but the finisher frame produces only about 60% of this because of the
reduction in sliver weight. Tow-to-top sliver making systems have replaced the
preparation up to and including the card in some cases but these modified systems
are capable of dealing solely with man-made fibers. The yarn is usually spun from
sliver. (Note: 100 in ≈ 2.5 m, 450 lb/hr ≈ 200 kg/h, 200 lb/hr ≈ 91 kg/h, 10s worsted
≈ 90 tex, 300 grain/yd ≈ 21 ktex, 50 grain/yd ≈ 3.6 ktex.)
8.3.10 Twisting and doubling
Yarn ‘folding’ is almost universal for worsted and woolen yarns or blends with wool
[9]. Folding is otherwise known as twisting or doubling. (See also Chapter 3.) Longstaple fibers, when well drafted, display low fiber migration, and the peripheral
fibers can easily be peeled off; folding stabilizes these fibers and improves the yarn
characteristics. Weaving folded yarns is accomplished with fewer end-breaks and the
fabrics contain reduced numbers of faults. Two-for-one twisting enables much longer
lengths of twisted yarn to be produced without a knot or joint. This, too, improves
weaving performance and reduces twisting costs; there is now even less burling and
mending of the fabrics and this has helped to assure the penetration of two-for-one
twisting in the industry. (Burling is the removal of yarn faults from the fabric.)
8.4
The woolen system
8.4.1 General comment
The woolen system often uses blends of fibers (sometimes with waste fibers) to make
a relatively inexpensive, soft, hairy, and full yarn. Fibers used are usually short by
wool standards and so are likely to felt. Fabrics made from woolen yarns are often
deliberately milled or felted to make them dense with a napped appearance on the
surface. Frequently, the yarn strength is low.
Layout of the processes is quite variable in this industry because of the diversity
of raw materials. Even if all the components are wool, the fiber properties might vary
widely and so might their state of cleanliness. To achieve consistency, it is necessary
to blend thoroughly before carding and this might involve several stages. For example,
a diversity of fibers might require that some fibers should be cleaned differently from
others and each lot might be blended within itself. Even after scouring, color differences
might be recognized that require further blending to even out the variations. The
intermediate product resulting from the preparatory and blending process is known
as a ‘willeyed blend’ and this product is the feed for the carding process. The card
Long-staple spinning
221
produces a condensed slubbing that goes direct to spinning. Woolen spinning has a
short and efficient process line that is less complex mechanically than some of the
other systems described. Since the process after the card feed is such a short one and
the feed materials are so diverse, there is little further chance of rectifying unevenness
(unless the material is reworked). Every opportunity is taken to blend the components
because any reworking degrades the material and adds unnecessary expense.
8.4.2 Woolen opening
According to Richards and Sykes [10], wool can vary greatly in its state. The conditions
range from clean to greasy and different amounts of dirt are sometimes embedded in
the grease. The dirt may be only loosely associated with the fibers. The size of fiber
clump can vary from locks to relatively large pieces. Thus, the amount of cleaning
and/or opening needed for the different categories varies widely. Excess opening can
damage or entangle the fibers and therefore it is sometimes necessary to process
different lots of fiber by different equipment running under different conditions
before any final blending.
The process of opening uses equipment not dissimilar to the types described for
ginning and opening in short-staple spinning or in worsted spinning. Compared to the
short-staple devices, the teeth or spikes used are larger but that would be expected in
view of the increase in fiber length. Suffice it to discuss only the differences from
types already described. The teazer (otherwise known as a wool willow, wool opener,
or devil) has many similarities to the opening machines already described. A significant
difference is that material is fed to the machine and is left for a predetermined time
(the ‘draw’) to undergo opening within the machine before being removed. This is in
contrast to the continuous flow systems described earlier. Another feature, not met
with elsewhere, is the possibility of grease build-ups in the equipment. The picker
opener is similar to machines already described and will not be further discussed. The
Fearnought opener is akin to a coarse-toothed card because it has a cylinder, workers,
and strippers. The teeth or pins are much larger than those found on cards. It is a
thorough opening device and promotes blending but it is never used to disintegrate
the wool pieces to single fibers or even to small tufts because this would interfere
with carding.
If recycled material is used, yarn or fabric has to be decomposed into fibers.
Briefly, the rags are shredded in a rag picker and then reduced further in a garnet to
separate the fibers sufficiently for them to be carded. Naturally this process is rather
severe and fiber breakage causes a considerable loss in fiber strength and may cause
damage that will show up later. Thus, although it is acceptable for the class of product
involved, there are hazards that are not usually found elsewhere. Yarns that have not
been completely decomposed into fibers cause trouble in carding and subsequent
processes. The so-called threads degrade the product. Some idea of the damage that
can be done to recycled fibers is illustrated by Fig. 8.8 where it will be seen that close
settings can severely shorten the threads, and by deduction, also the fibers. Recycled
material is no longer widely used because of governmental regulations regarding the
labeling of reused fiber: the introduction of many different blends and types of fiber
has made the identification of the fibers in the rags difficult and this makes compliance
with regulations difficult. Furthermore a great deal of the economic motivation has
disappeared, and therefore the traditional woolen system is becoming rare. In modern
times, large volumes of synthetic fibers are used in the blends. The opening process
222
Handbook of yarn production
300 in/min
Thread length (inches)
3
64s wool
150 in/min
2
300 in/min
1
50s wool
150 in/min
0
0.015
0.020
0.025
Breaker setting (inches)
Fig. 8.8 Thread length
has been replaced by more conventional fiber opening and cleaning, using virgin or
man-made fiber. Where virgin wools are used, layers of fiber are oiled to plasticize
and lubricate them. The man-made fibers need little opening and cleaning; consequently
they are introduced later along the process line.
8.4.3 Woolen blending
Hopper feeders are still used to blend fibers. The opportunity is often taken to stockdye fibers. In such a case, the material then has to pass through squeeze rolls, a dryer,
and an oiling section before further processing. After opening and cleaning the fibers
to bring them to a compatible state, the main process of blending occurs. Sometimes
this is manual but, increasingly, mechanical systems are being used.
In blending, good stock records and good housekeeping are prerequisites to
satisfactorily uniform blends. Mistakes at this point usually involve reworking the
material, which (a) is an unnecessary expense, (b) increases fiber damage, and (c)
adversely affects carding. When reworking becomes unavoidable, it is preferable to
take the material from the line before it reaches the Fearnought process, to avoid as
much fiber damage as possible. With manual blending, human variability gives rise
to additional error possibilities. A reasoned procedure, which takes into account the
varying properties of the feed material and the needs of the product to be made,
assumes an even greater importance. In manual blending, the components are spread
as layers, in an order determined by the specified plan, to take into account mass and
composition of the various feed lots. The thickness of each layer has to be as small
as is consistent with an even coverage of the area of the laydown. Sometimes the feed
lots are pre-blended by various means. Obviously this is a labor intensive and skilled
operation. In a climate of high wages, mechanical devices will eventually replace the
traditional methods. Modern methods favor pneumatic handling of the fibers and
mechanical means of layering or mixing. One form of mechanical blending is the
rotary spreader, which deposits a circular layer of fiber in a suitable bin or receptacle.
Several feeds can supply the spreader at the same time. An alternative is to pre-blend
intermediate lots into the final blend. If Z sections of intermediate lots are each built
from the same feed lots and in the same proportions, then a blend with good longterm evenness will result. No matter what the magnitude of Z is, long-term differences
are thus avoided. Various opening type machines can be used to improve the local
blend, providing they do not break down the tufts too much or damage the fibers. It
Long-staple spinning
223
is possible to supply the willeyed blend directly to a Fearnought or other blending
machine if the production rates are properly synchronized. Such a line can be integrated
with an oiling system. A site to add the processing agent to the wool is needed.
Scouring removes the natural lubricant.
8.4.4 Willeyed blends
A customary final blending is to pass the material through a Fearnought willey that
blends what should be small wool pieces or tufts. Despite the care in preparation,
willeyed blends still contain differences. Cleanliness can vary, especially if scoured
wool is bought and added to the blends. The wool can contain varying amounts of
coloration, processing lubricant, burrs, and other vegetable matter. Where man-made
fibers are to be blended, no cleaning is needed and the materials are blended at the
latest possible stage. Such blending is almost an irreversible process where separation
of the fibers becomes quite impractical and the possibilities of reworking become
correspondingly smaller. The question of what lubrication is needed for such blends
is heavily dependent on the fiber finish applied by the man-made fiber maker.
Inhomogeneity arises if the wool pieces are large or the building of lots has been
irregular. Sampling and testing inhomogeneous feed lots and the resultant blends are
important parts of the technology, the variability; and characters of the feed lots
determine sampling rates.
The material is often baled and stored until required for use in further processing.
Alternatively, it is stored in large bins, in which case pneumatic fiber transfer systems
can be used. The material within each of the bales or storage bins should be as nearly
homogeneous as possible in tuft size, fiber composition, and cleanliness. Of course,
if stock-dyeing has been used, the variety of dye shades increases the difficulty of
stock and quality control.
8.4.5 Initial woolen carding
Conventional woolen cards are fed by card hoppers that usually have a weighpan
device to control the mass flow of fibers. Fiber is carried from the main storage bins
by lattice feeds to the hopper. The action of the hoppers removes much of the variability
that comes from manual feeding. A hopper feeder has two or more bins as part of its
structure, together with the necessary moving lattices. Fiber from the large bin is
carried on an inclined lattice, which supplies fiber to the smaller weigh-bin. Rakes
are used to roughly level the fiber sheet on this lattice and the fibers, which are
removed tumble back into the bin. The tumbling action gives a degree of blending
and the action of the pins opens the fiber masses somewhat. However, the tumbling
can cause aggregation of fiber clumps to unacceptably large new ones. In such cases,
the danger of choking becomes greater. The solution is to use appropriate lattice
speeds so that the excess removed is not too large. Controllers are used for the lattice
drivers and this improves the accuracy of control, especially at high production rates.
The contents of the weigh-bins are dumped periodically (several dumps/minute)
on another moving lattice and the flow control is based on the mass of fiber in each
weigh-bin. Often the dumping rate is partly determined by the lattice speed. There
can be several hoppers, all of which dump their discharges on a common lattice feed,
which goes to the card to give another stage of blending: in this case, the dumping
times of the various hoppers are co-ordinated. Devices using strain gages and/or
224
Handbook of yarn production
other sensors are displacing mechanical weighing systems, and the electrical signal
outputs facilitate the use of computer control. Such computer control is capable of
leveling several consecutive weighings to deliver a moving average of the input fiber
mass and this reduces long-term irregularity in the slubbings. There is a trend to the
use of chute feed systems, often in co-operation with a type of hopper feed; this
reduces the need for labor and might provide a site for control.
Since the card feed is one of the few practical points for mass flow control, further
measurements of the feed sheet thickness are made and there are a number of methods
available to do this. The use of a dancing roller, with sensors to detect changes in
height, is one way that gives an average across the width of the sheet. Measurement
of the penetration by ultrasonic or gamma beams through the sheet provides an
alternative source of control signal. In these cases, it is necessary to either (a) choose
measurement sites that sample the sheet accurately across the width, or (b) use a
scanning device which moves to and fro across the width of the sheet. One reason for
the extra control stage is that changes within the hopper can produce errors, especially
if there is a tendency for the feed system to choke. Evenness also is needed in fiber
proportions and state. The action of pinned rolls can not only fractionate the material
delivered into different fiber types and clump sizes, but also cause agglomeration of
smaller clumps into larger ones [10]. Varying quantities of dirt, oils, residual grease,
and other materials can influence performance and it is quite likely that the measurements
used to control the fiber flow will not properly reflect these causes of error. Moisture
and some processing agents evaporate over time and this is another cause of variation,
especially if water sprays are used in blending. Careful human monitoring is needed
to ensure that the blend is within the limits of uniformity in all respects.
8.4.6 First stage in woolen carding
Woolen cards (Fig. 8.9), like worsted ones, nearly always have roller tops because
they are less likely to become clogged than non-rotating elements. Locks of fiber can
be handled with relative ease by the worker/stripper in combination with a swift (i.e.
cylinder). It may be recalled that the workers and strippers perform a blending function
because fiber is not immediately removed from the worker/stripper sub-system. The
lag in fiber transfer acts to mix early arriving fibers with later ones, as well as to
perform the fiber opening and orientation actions of the sub-system.
To give some idea of the dimensions of these cards, a typical swift is about 5 ft
(≈ 1.5 m) diameter, the workers and strippers about 8 inches (≈ 0.2 m) and 4 inches
(≈ 0.1 m), respectively, and the whole card set might be 20 yd (≈ 18 m) long. The
width can be several yards (meters) and the swifts might rotate at up to 200 r/min.
These card sets are enormous pieces of machinery and they can process material at
speeds up to 500 lb/hr (≈ 227 kg/h), although the range of speeds and sizes varies
considerably. Carding forms a large part of the processing set-up in woolen spinning. This
means that if a card set becomes non-operational, a large fraction of the production is lost.
Woolen carding is noted for the wide range of equipment groupings and the best
that can be done in limited space is to describe a typical set-up. Such a set-up might
include a feed system, a breast with perhaps three workers and strippers, a scribbler
with perhaps four sets of workers and strippers, a cross lapper, a two-swift carder (i.e.
two-cylinder card) with a total of about eight workers and strippers, crush rolls, and
a condenser that produces the slubbings. Let us start the description at the feed end.
A silhouette of a typical first unit of a set is given in Fig. 8.9.
Long-staple spinning
Worker
225
Stripper
Fancy
Swift
Breast
Delivery
Feed
Transfer roll
Fig. 8.9 Typical first carding unit of a woolen system
Weighpan hoppers (not shown) drop fiber onto a moving lattice that forms the feed
to the card set. The feed rolls are often pinned to grip the fibers without undue
compression. The items shown in dark gray may have fairly coarse garnet clothing,
whereas later sections may have flexible wire. Garnet clothing is a metallic wire
similar to that used in short-staple processing, but with larger teeth. The delivery
from this unit may pass to a scribbler (breaker card), which has a similar swift/
worker/stripper arrangement, although the swifts are usually larger. The delivery
from the last swift of the first carding unit is doffed and passes through a crush roll
set. At this point, embrittled vegetable matter is reduced to powder and drops out of
the web as it passes through to the carder by way of a cross-lapper.
8.4.7 Cross-lapping
The woolen system differs from the worsted in several different ways. One of the
most important differences is that the output often consists of a number of slubbings,
each drawn from a narrow band of fiber from the card web. A slubbing might be
taken from a region ranging from one selvage to the other. It is highly undesirable for
the material to vary over that range otherwise the slubbings would be dissimilar. It is
therefore essential that the web be even across its width. One way of improving the
across-width evenness is to cross-lap, using a Scotch feed or similar device, somewhere
between two of the main cylinders (usually between the scribbler and the carder). The
principle is illustrated in Fig. 8.10. Web enters at W on a feed lattice oriented
perpendicular to the paper and oscillating in the direction X. The web is laid zigzag
fashion on another moving lattice, which moves in direction Z. The direction of
oscillation X is usually roughly perpendicular to the direction Z. In many practical
set-ups there are two such cross-lapping sections. Web-feed lattices are arranged to
complete the changes in orientation of the body of the sheet, so that the feed to the
next machine is again in line with the original flow.
8.4.8 The carder
The cross-lapped web then passes to the carder (finisher card), a diagrammatic
arrangement being shown in Fig. 8.11. As before, each cylinder has its quota of
workers and strippers and the theme of blending continues. It has to be mentioned
that many stages of division of fiber tufts are required. Interactions between the
swifts, strippers, and workers throughout the system cause tufts to be broken down
into smaller ones and thence (mostly) to single fibers. Fortunately, there is incomplete
226
Handbook of yarn production
W
X
S
Y
Z
Fig. 8.10
Cross-lapping
Workers and strippers
Fancy
Fancy
Swift
Swift
Doffer
To tape condenser
Transfer rolls
From X lapper
Fig. 8.11
Carder
doffing at these various interaction points and a typical fiber lock recirculates many
times; the fiber receives several hundred stages of opening and blending before it
leaves the system. The carder delivers web to the tape condenser and all these major
units are set in line to make a continuous system.
Dirt gathered under the card has to be removed and safety regulations require the
machine to be enclosed. Manual cleaning requires a shutdown if a risky cleaning
operation is to be avoided; consequently pneumatic systems are likely to become
more prominent. Management of the carding system takes skill and experience.
Flexible wire has not been discussed elsewhere because it is obsolete in shortstaple spinning, but it still serves an important function in woolen spinning. In working
wool, the fine wires are effective in teasing fibers apart without undue damage. The
Long-staple spinning
227
wire population density varies according to use but it ranges from 100 to 800 points/
inch2 (≈ 0.15 to 1.24 points/mm2). Wire, usually of round cross-section, is embedded
in a base material that is mounted on the surface of the element concerned (e.g. swift,
worker, doffer). The cross-sectional area of the wire is small compared to its length
and this is what gives it its flexibility. The shape is cranked as shown in Fig. 8.12(a)
because it is necessary to have an adequate angle of attack, α, and the angle, β, is
chosen so that the height does not change under load. The clearance between cooperating wire surfaces might be as little as 0.008 inches (≈ 0.2 mm) and there is little
tolerance for reduction. Changes in the height of the wire alter the setting, and if the
setting is reduced, very expensive damage can occur.
Properly selected and maintained wire is important. Another problem associated
with flexible clothing is its tendency to load. Fiber becomes embedded deep into the
wire and, after a time, this can impair the efficacy of the card. For this reason it is
necessary to use a ‘fancy’ to bring fibers to the surface. The fancy is made of long,
flexible wire, and the wires penetrate those of the swift. Despite the fancy, the wire
will still load in time. Periodic stripping or fettling of the wire is necessary, which
means that fiber trapped by the card wire is removed. Depending on the fiber being
processed, it is possible to card between 1000 and 8000 pounds (≈ 450 to 3600 kg)
of fiber between fettlings. A card behaves abnormally after fettling until a sufficiency
of fiber has been deposited in the wire to give equilibrium conditions. Thus, maintenance
involves not only the grinding of the wire, but also the cleaning of the wire interstices
(i.e. fettling) and re-establishment of equilibrium conditions. An overloaded surface
will not collect fibers and the process begins to break down. As the loading increases
beyond a certain limit, the weight of the web diminishes and so does the eventual
yarn count. Loading of the cylinders also affects the opening power of the card and
it has to be controlled to produce a good yarn without slubs or defects.
The cylinder, worker, and stripper speeds have to be adjusted to suit the fibers
being processed. Excessive worker and stripper speeds damage the fiber because of
the increased rates at which the fiber tufts are drafted. Workers on the finisher swift
(cylinder) should be limited in speed to prevent irregularities, but those on earlier
α
103
(i)
64s wool
(ii)
Nep/g
102
10
(i) 300 in/min
(ii) 150 in/min
50s wool
(i)
(ii)
β
1
0.015
0.020
Setting (inches)
(b)
(a)
Fig. 8.12
Flexible wire and nep
0.025
228
Handbook of yarn production
cylinders might be speeded up to increase the opening effect. Doffer speed and feed
rates affect the loading, as do the settings, and these affect performance. Production
rates depend on the speed of the swift. Settings and speeds become critical.
The mechanism of transferring fibers calls for comment. The surface speed of the
swift is much faster than that of the worker or doffer, and the fibers on the web
surfaces are layered like overlapping tiles, with straightened ends projecting forward.
Portions of the fibers on the doffing surfaces are brushed by the others on the swift
to produce fiber hooks, which affect subsequent processing. The size of the doffer
has little effect on the fiber transfer rate, providing the surface speed is kept constant.
Moreover, the direction of doffer rotation does not affect it either [6].
As with short-staple spinning, the condition of the wire is important, as are the
settings. Worn wire leads to high nep production, which has just as deleterious an
effect on quality as in the other systems. The settings are somewhat more difficult to
maintain because the wire is not always as rigid as with the systems previously
discussed. A definite trend for increased nep was shown as the settings were increased
in the experiment portrayed in Fig. 8.12(b). The designation of the wool as an Xs
means that the wool is suitable for spinning to an X count of yarn. Thus a 64s wool
should be capable of being spun to a 64s yarn. It is a designation of the expected spin
limit.
The web leaving the doffer presents another of the easily accessed control points.
The web thickness can be measured by various means and the doffer speed can be
controlled using the signal from the transducer. Photoelectric web thickness measuring
equipment used to adjust the doffer speed can be used to improve the productivity;
this is called the autocount system. There is a considerable inertia effect and no shortterm control is possible. Limits on the loadings on the swift mean that the average
speed of the doffer has to be coupled with the input to the system. Very long-term
errors cannot be controlled from signals obtained solely from signals derived from
the web; long-term control is better done at an earlier stage. To obtain more accurate
control, signals can be obtained from intermediate positions by measuring the fiber
on the swifts or doffers using scanning, flat laser beams, or other devices. There is a
band of error wavelengths that can be controlled from signals generated by transducers
on the card, providing that the long-term average mass flow is also controlled. The
web is still a major source of error in woolen yarns. Apart from linear density, longterm variations in the amount of oiling or opening or blending adversely affect the
operation; fiber loss or degradation varies accordingly. The blending length of the
card set, although very long, is still limited, and the card set cannot remove all these
errors. A blending length is defined as the length of web over which variations will
be attenuated to a prescribed level. A short blending length implies that the integration
process in the card merely ‘smears’ the variation over a short length and even longerterm variabilities pass through almost unsmoothed. The woolen card set is deliberately
designed to have a very long blending length. This is because all irregularities in
composition and thickness of the web are passed directly to the slubbing and may
show up as variations in linear density, color or some other attribute. Thus, every
opportunity has to be taken to improve evenness.
One of the several purposes of the card set is to remove unwanted material and the
machine is quite efficient at doing that. If, however, the percentage of unwanted
matter varies widely within the blend, its removal produces a complementary irregularity
in the fibers delivered to the tape condenser (see following section). If the feedstock
is properly prepared, this is not too much of a problem, but unless experienced, the
Long-staple spinning
229
workers in preparation might not realize the importance of their work. In addition, the
card set can still contribute its errors to the products. Errors produced in, or near, the
tape condenser show up as relatively short-term errors in the slubbings. Consequently,
there is a very wide spectrum of error and there is no drawframe doubling to even
them out. Thus, it has to be again emphasized that there needs to be a great deal of
attention to evenness at all stages.
8.4.9 The tape condenser
A prominent difference between the conventional woolen and worsted systems lies in
the card delivery. Traditionally, a woolen card set delivers slubbings rather than
sliver, and the slubbings are converted to yarn in a spinning machine. However, some
woolen systems produce sliver, much of which is used for sliver knitting. The system
is also used for non-wovens. These last two uses do not involve yarn and will not be
further discussed.
The slubbings or ropings are like roving but have no recognizable twist. A woolen
card delivers up to 200 such slubbings, which are created by splitting the web into a
number of similar ribbons of web that are rubbed to give the needed fiber cohesion.
Consolidation of the fibers at this stage helps retain the shorter ones and avoid
undue fiber loss; the number of active tapes is related to the yarn count required and
the web thickness. The system of separation of the sheet into tapes is designed to
limit fiber damage, but inevitably there is some short-term irregularity caused by the
process. Card web is laid onto a series of tapes, which separate the ribbons (Fig.
8.13). There are a variety of tape arrangements possible but the figure of eight pattern
is probably the most popular. The tapes have to be durable and retain their surface
characteristics over long periods of time. Slick or greasy spots on a tape can cause
local irregularities in the web being divided and so can wear at the edges. Variation
in tape tension is another factor, particularly between banks of condensers. The web
selvages are usually too uneven or of the wrong weight and are usually discarded or
reworked. The useful ribbons are about 1 inch (25 mm) wide and they pass to pairs
of rubbing aprons that roll them into ropings. The whole assembly is called a tape
condenser. Each slubbing is taken up on a spool and these go direct to spinning. The
Slubbing
output to
winding
section
Reciprocating
rubbing aprons
Web division
Twisted tapes
(b)
Card
web
input
Tapes
(a)
Fig. 8.13
Woolen condenser with rub aprons
230
Handbook of yarn production
material at this stage is fragile and needs careful handling. Out-of-true carding elements,
periodic differences in loading, wire sharpness, etc., produce periodic error in the
slubbings that can be measured in the laboratory. Truly periodic errors can be recognized
and the source of error can usually be determined. However, there is often a larger
and more distributed random component for which diagnosis is difficult. Experience
and trial-and-error tests are needed to find the sources of these types of errors. Of
course, with such a wide range of fiber properties, high levels of random errors are
to be expected.
A card set stores a large volume of fibers during operation because of the size of
the elements involved and the degree of loading. This has advantages in averaging
medium-term errors but it causes significant inertia in the system, which is particularly
noticeable on start-up. As the fiber delivery rate alters, so does the wire loading; it
moves exponentially to some steady value. Stability is not achieved for several minutes.
Care has to be taken in disposing of the slubbings produced during start-up or shutdown
procedures. For a five-part card, the transient can be as high as 8 or 9 minutes.
8.4.10 Woolen spinning
Slubbings are usually formed into groups of cheeses on a mandrel and the assembly
is referred to as a bobbin. These are then transported directly to the spinning machine,
which is often in the form of a ring spinning machine. (Mule spinning still survives
in some areas but space precludes a discussion of this.) Bobbins often have flanges
to protect the material, because the cheeses are soft and are vulnerable to damage.
Denser packages are preferable, because of the more efficient use of space, and it
then becomes possible to dispense with the flanges on the bobbins. However, the
denseness of the cheeses is quite dependent on the tension that the slubbing can bear,
as well as on the wind structure. Undue tension can cause uncontrolled drafting and
end-breaks; this represents a limit to the possibilities. Winding lag has to be avoided
to get the improved package density that arises from a precise lay. One solution is to
use a grooved tension plate. The objective is to lay the material directly on the surface
of the cheese, near the nip, between the spool and a drive roll.
Emerging practice is to transport a series of bobbins by endless chains, and creel
them in the spinning frame automatically. Arrangements have to be made to ensure
that a given lot of slubbing arrives at the proper spinning frame and obviously computer
control has a part to play in this. Direct spinning from the slubber has not found favor
over the years, probably because of the lack of flexibility inherent in such a system.
In spinning, a strand’s weakness makes necessary special precautions in order to
locally strengthen the yarn at the weak points. In drafting, false twist is introduced to
bring the drafting point closer to the nip of the delivery rolls than otherwise would be
possible with large rolls. The false twist runs to the twister surface. Consequently, the
effective ratch setting is between the twist transition point and the nip of the delivery
rolls. A typical system is shown in Fig. 8.14. The false twist spindle runs at anywhere
from 20% to 60% of the main spindle speed. Higher percentages apply mostly to the
higher speed spindles [10]. The lower part of the drafting section usually comprises
a series of rolls that are fairly conventional. Fiber control is also needed in drafting.
Fig. 8.14(b) shows a typical set of rolls. Rolls A and B are the ones shown in diagram
(a). D is a fiber control device that fulfills much the same function as an apron in
short-staple spinning. The distance between D and E is a function of the fiber length.
The output is twisted and wound by a ring spinning machine similar to those
Long-staple spinning
231
Back rolls
False twist draft zone
Roping
From the false twist draft zone
False twister
A
B
Front rolls
D
Main draft zone
C
E
Yarn
To ring bobbin
(a)
(b)
Fig. 8.14
Control of fibers in woolen drafting
already discussed. The tensions caused by spinning are reduced by using collapsed
balloons or balloon control rings. Woolen yarns are used as both singles and ply
yarns. In this trade, plying is known as doubling. Since woolen yarns have the
characteristic of being soft, with good insulation properties, singles yarn has a market
for hand knitting yarns and garments where those qualities are prized.
8.5
Bast fiber spinning processes
8.5.1 Conversion of stems to sliver
The process of converting flax, jute, and hemp consists of hackling, preparation, and
spinning. The first step (i.e. hackling) is to (a) split and separate the fibers that are
gummed together at the start of the process, (b) disentangle them, and (c) parallelize
them as far as possible. The remaining broken, shorter, raveled fibers form a tow,
which is a byproduct of the process and regarded as inferior. Yarns are also made
from this tow. Traditional hackling was performed by hand, using spiked boards, but
more modern practice uses machinery. The first stage of hackling is known as roughing.
A principal component of a roughing machine is a moving band carrying spiked bars
containing hackling stocks, which work on the ‘stricks’ of rough flax and carry out
the first rough separations. Successive stages have ever finer teeth, more closely
packed to finish the process, the finest pitch being about 60 pins/inch (≈ 2.4/mm).
Fiber bundles are moved from one hackling band to a neighboring finer-toothed one
for further processing. This process continues until the bundle reaches the finest
hackling band. The root ends are hackled first, the ‘combed’ end is clamped, and the
other end is hackled in a manner similar to that just described. The tow made during
these processes is stripped from the teeth of the bands by brushes whose surface
speed exceeds that of the pins. The stricks of hackled fiber are then sorted into
different qualities; smoothness, luster, hand, and cleanliness are factors that determine
the quality.
232
Handbook of yarn production
Fiber bundles of similar quality are fed to spreading frames which transport sheets
of fibers to a drawfame for a gilling operation similar to that already described. The
lengths of flax on the spread sheet entering the drawframe must overlap to give
cohesion and evenness; also, several parallel sheets are used to give a doubling that
reduces unevenness. The output of the drawframe, just as in the other similar processes
described elsewhere, is referred to as sliver and is stored in cans. In all, it is practice
to use at least four drawings and to creel between four and twelve slivers to give a
large amount of doubling. It is common to use a cumulative doubling ratio approaching
1500:1. (It will be remembered that the doubling ratios for each stage are multiplied
together to calculate the cumulative doubling ratio.)
Manufacture of yarn from tow follows a different process. First, the tow contains
shives (woody material) and other impurities, which have to be removed. Opening
and cleaning equipment is somewhat similar to that used in short-staple spinning. It
is then carded using one- or two-card sets, depending on the quality and cleanliness
of the input material. The card sliver is then drawn and spun in similar fashion to that
used to produce other bast fiber yarns.
8.5.2 Spinning bast fibers
Spinning can be done using one of two systems. The dry spinning system is used for
coarse yarns; a roving stage similar to those already described is used to produce an
intermediate product. Sometimes another flyer spinning system is used to produce
the final yarn and sometimes a ring frame is used. Wet spinning is used for finer
yarns. The rove (roving) is passed through a trough containing hot water and the rest
of the spinning is carried out wet. The water dissolves the gummy substances and
provides freedom for the fiber to slide in a controlled fashion in drafting, with the
result that evenness is much improved. Twistless spinning of cotton using wet drafting
showed the same effects [11]. Spun yarn is then usually wound into hanks containing
300 yd. These hanks are referred to as leas or cuts. The grist (count) is calculated by
the number of leas/lb.3 There are other count systems, which will not be enumerated.
The hanks are dried and then worked by twisting and untwisting them to dispel the
wiry feel of the yarn; this breaks down the gummy adhesions, which give the wiry
hand. Fine linen yarns are often bleached before passing to the lace maker or weaver.
Linen thread involves the plying of very fine linen yarns.
References
1.
2.
3.
4.
5.
6.
7.
Wood, G F. Wool Scouring, Text Prog, 12, 1, 1982.
Von Berger, W. Wool Handbook, Interscience, New York, 1982.
Atkinson, K R and Saunders, R J. Burr Beater Design and Operation, Part 1, J Text Inst, 82,
4, 1991.
Plate, D E A. Advances in Early Stage Processing of Wool, Textile Inst Conf Proc, Sydney,
Australia, 1988.
Atkinson, K R. An Analysis and Theory of Burr-beater Operation, J Text Inst, 80, 2, 1989.
Robinson, G A. High-speed Carding of Wool, J Text Inst, 80, 1, 1989.
Anon. Processing of Staple Fiber on Woollen and Worsted Systems, Technical Booklet 12,
ICI, Harrogate, UK 1970.
3 This is yet another yarn count system to add to those described in Appendix 1.
Long-staple spinning
8.
9.
10.
11.
233
Henshaw, D E. Worsted Spinning, Text Prog, 11, 2, 1981.
Lorenz, R R C. Yarn Twisting, Text Prog, 16, 1/2, 1987.
Richards, R T D and Sykes, A B. Woollen Yarn Manufacture, Manual of Textile Technology,
Textile Institute, Manchester, UK, 1994.
Selling, H J. Twistless Yarns, Merrow Monograph, Watford, UK, 1971.
9
Post-spinning processes
9.1
Winding
9.1.1 Introduction
Bobbins from a ring frame contain too little yarn to be useful in modern fabric
making equipment and it is necessary to rewind the yarn onto larger packages.
Special, high speed winding machines are used for this purpose. It is very important
that the yarn which is to be sold or used in a subsequent process should be ‘put up’
into the correct sort of package – usually cones, cheeses, or occasionally, hanks. The
transfer of yarn from the ring tube to the cheese or cone provides an opportunity to
remove yarn faults. Looking towards the consumer’s needs, it has to be realized that
the density and structure of the package delivered are important. For transport and
storage, the package should be as dense as possible. For ease in unwinding, the
package should have a regular structure without over-dense portions, which might
impede the unwinding process. Poor unwinding properties cause difficulties for the
user and increase the costs. For dyeing, a low but regular package density is required
so that the dye liquor can penetrate the package easily and evenly. Irregular dye
penetration yields streaks and barré in the final product. Variations in winding tension
produce similar effects. The needs of the customer or user therefore dictate the type
of package and the density of winding. If the yarn is returned as a complaint, the
spinner’s costs are increased.
Dye packages are usually wound on sprigs (porous package centers) and are shipped
mounted on pegs, which form part of a transport frame. Sprigs permit the flow of dye
liquor in dyeing and the peg-frames prevent the packages rubbing together and becoming
damaged.
As previously stated, ring frames produce low volume bobbins that contain blemished
yarn; also, these bobbins have a combination wind, which is unsuitable for the next
process. The final yarn packages are usually cross-wound and contain several pounds
of yarn. This means that there must be many joins in the yarn on each package
because there must be at least one join for every ring bobbin used. To these must be
added another one for every blemish removed in the clearing operation. Yarn faults
Post-spinning processes
235
outside the prescribed limits are removed as the yarn is transferred from the spinning
bobbins to the cones or cheeses. Faulty portions of yarn are cut out and the ends are
spliced together to make, as nearly as possible, a perfect join; all of this is done
automatically. This latter process is called ‘clearing’ and, while it is very effective, it
is better to have as few a number of original yarn faults as possible, rather than rely
on the clearing capabilities of the winding machine. Each intervention by the winder
adds to the cost and slightly degrades the quality of the yarn. Winding and clearing
of staple yarns are normally carried out on the same machine.
In filament production, and in certain advanced staple spinning systems (such as
rotor spinning), the primary process produces a large, cross-wound package and there
is no need to rewind to change the package size, although sometimes there may be a
need to rewind to a low density package for dyeing and sometimes there may be a
need to clear defects. However, every effort is made to avoid such costly rewinding.
The technology of winding has developed to the extent that automation is the rule.
It is therefore important to consider this aspect fully. Economics and quality control
become very important factors in the seemingly simple task of rewinding yarn (or
‘winding’ as it is normally called). Winding is carried out for the following purposes:
1
2
3
4
To
To
To
To
change the type of wind.
change the package density.
remove yarn faults.
create a package which is not susceptible to damage.
9.1.2 Machine principles
Machines are required to produce acceptable yarn packages as just outlined; this
often involves the manufacture of a cross-wound package using a reciprocating guide,
as shown in Fig. 9.1. Consequently this type of machine will be used to open the
discussion. The yarn is laid on the surface of the rotating package by this reciprocating
guide. The idea is that overlapping sinusoidal wraps of yarn interlock and provide a
stable package. The relative rates of traverse of the guide and yarn package determine
the type of build. As the package grows in size, its tangential speed increases unless
it is controlled in some way. This implies that, for a given traverse oscillation rate, the
geometry of each layer of yarn gradually changes as the package grows. As will be
seen later, this has some important consequences.
9.1.3 Package build
The package build most used is a cross-wind in which the yarn is traversed across the
face of the cheese (or cone) several times during one rotation of the package. For
convenience, a cone or a cheese will be referred to as a package because many of the
following remarks apply to both.
Yarn on the surface of the package is roughly sinusoidal and, ideally, out of phase
with the coils lying beneath. Yarns cross one another and friction holds most of the
yarn in an × formation (Fig. 9.2(b)) which is highly stable; this makes packages very
durable under normal conditions if abuse in handling is avoided. The winding tension,
angle of wind, and cradle pressure affect the structure and package density. The socalled ‘cradle pressure’ refers to the force acting between the package and the drive
roll. The density of a regular package should be controlled because it might later be
236
Handbook of yarn production
Package
Yarn
Reciprocating guides
(a)
Package
Yarn
(b)
Fig. 9.1 Cross-wound package
unwound at high speed and too hard a package then gives trouble.Winding tensions
should be limited due to the danger of overstraining the yarn. Changes in cradle
pressure cause changes in winding tension and adjustment of the cradle pressure is a
means of exerting control on tension.
As the diameter of the package builds up, the number of traverses per rotation of
the package changes and this changes the structure of the package. When the package
is the same diameter as the drive roll, the coils on the surface of the package lie
exactly on top of the ones just laid. A number of coils laid on top of each other as the
diameter builds through the critical zone produces so-called ribbons that are very
troublesome (see Section 9.1.4). In the case of a dye package, the dye penetrates the
ribbon at a different rate from the rest of the package; the result is a periodic difference
in dye shade in the yarn. In unwinding, the yarn is reluctant to leave the surface of the
dense parts and this gives rise to tension pulses which also can cause difficulty.
Ribbon breakers are normally fitted to the machines to prevent the problem. Some
machines cause the package to lift from the drive roll momentarily, to allow some
slippage to disperse the ribbons at the crucial times. Other machines move the package
sideways to achieve the desired dispersion. Some modify the drive roll speeds.
The yarn in the shoulders of the package plays a considerable role in the durability
of the package. There should be sufficient traverses per revolution of the package to
prevent loose portions of yarn from lying parallel to the shoulders. The shoulders
should feel firm and stable, but not hard.
As has been explained earlier, yarn is sometimes parallel wound. In such a case,
either the package must be wound onto a bobbin with flanges, or the package must
be small and have sloping shoulders. Hank or skein winding can also be used. Sometimes
skein dyeing, and occasionally skein mercerization, is used; skeins are often sold in
the home craft market and they frequently consist of heavy yarns.
Post-spinning processes
237
9.1.4 Cross-wound packages
For simplicity of explanation we will confine ourselves at first to a cheese. The
reason for stability can be seen in Figures 9.2 and 9.3, where it may be noticed that
the yarns on the surface of the package interlace at an angle. Each layer of yarn
imposes a restraint on the sinuous ‘coils’ beneath. The best angle is between 12° and
20°. Resolving the tensions in a given layer of yarn, the radial components acting
inwards at every intersection have a magnitude of F; these depend on the winding
tensions, T, and the angle of wind, θ. The number of intersections depends on the
wind, and θ again enters the picture. The radial component and the number of
intersections determine the density of the cheese, and the package density is a function
of T and θ as shown by the example in Fig. 9.3(b) and (c). The force needed to make
one coil slide over the others depends on the coefficient of friction µ, as well as F, and
the number of intersections, m. A force greater than µmF, acting along the length of
the yarn, could cause whole coils of yarn to slip. A cross-wound cheese has a large
number of intersections and it is, therefore, inherently stable. It is quite possible to
build a stable cross-wound cheese containing over 10 lb (4.5 kg) of yarn. However,
it still has to be remembered that the winding tension is an important factor in
determining both the package density (Fig. 9.3(c)) and stability. As previously mentioned,
too high a tension can damage the yarn; the aim is to maximize the stability without
exceeding the tension limit. The limit depends on the type of yarn being wound.
Referring back to Fig. 9.2, a phase change, φ, occurs as each layer is added and:
φ = πD – mλ
[9.1]
where m is an integer, λ is fixed and D is variable. As the cheese builds up, φ changes
periodically with the consequence that the package structure also changes periodically.
πD
λ
φ
C
A′
A
C′
B′
B
(a)
(b)
1
2
(c)
Fig. 9.2
Build-up of package surface
238
Handbook of yarn production
T1
F
θ
D
T2
(a)
Package density (g/cm3)
0.5
0
0
30
60
Angle of wind (θ degrees)
(b)
90
Cradle pressure = 2 lb (0.9 kg)
0.5 Package diameter = 6 in (0.15 m)
Ne = 20
0
0
Fig. 9.3
10
Winding tension (gf)
(c)
20
Cross-wound packages
At the times that φ is reduced to zero, yarn from one ‘layer’ is laid exactly upon the
one below and the yarn piles up in a dense sinuous ribbon on the surface of the cheese
until the diameter grows sufficiently to give φ a significant value again. As m reaches
successively larger integers, new ribbons are created; this causes considerable problems
in dyeing and unwinding. The flanks of the package also show signs of the ribboning.
At the critical diameters, the effective yarn traverse is increased and the reversal
points protrude; these are called overthrows (Fig. 9.4(c)).
Not only do patterns occur when φ passes through zero but they also occur when
φ is a fraction of π D (i.e. when φ/π D = 1/2, 1/3, 1/4, 3/2, 5/2, etc.). A solution to this
problem is to vary the lateral position of the yarn lay, as shown in Fig. 9.2(c).
There is an inherent variation in package density in a cross-wound package. Consider
two slices, A and B, of equal width, δl, taken perpendicularly to the axis of the
cheese, as shown in Fig. 9.4(a) and (b), where just one coil of yarn in the geometrically
developed surface is shown. Of course, each slice contains many yarns. A typical
length of yarn in slice A is δlA and it lies parallel to the shoulder. In fact, all the yarns
in the slice are more or less parallel. The lengths in slice B, such as δlB, criss-cross
Post-spinning processes
239
(a)
B
A
δl
δ lA
A
B
δl
δ lB
(b)
(c)
Overthrows shown by arrows
Fig. 9.4 Hard shoulders and overlaps
within it. Thus, the shoulders of the package at slice A are denser than those within
slice B or any other intermediate slice. Therefore there is a need for several traverse
motions per revolution of the cheese to keep the percentage of yarns lying more or
less parallel to the shoulder.
Ribboning and dense shoulders also occur with cones and although the mathematics
are more complicated, the ideas are essentially the same. Ribbons occur periodically
with the same unwanted effects, and the solutions are similar to those already described.
A minimum of several traverses per revolution of the cone is also required.
9.1.5 Cones
Although a cheese is easier to wind than a cone, there are advantages that favor the
cone. In knitting, yarn is withdrawn slowly and there is insufficient speed to cause the
yarn to balloon away from the surface of the cone. Yarn drags over the surface of a
cheese and disturbs the lay of the other yarns; the drag generates more and variable
yarn tension, as well as making the yarn more hairy. With a cone, the taper causes a
progressive release of the yarn from the surface, providing it is withdrawn in the
direction of the apex of the cone. There are fewer surface entanglements and the yarn
flows more evenly. Even at the higher speeds used in warping and high speed weft
insertion, the cleanliness with which the yarn is withdrawn from a cone is a considerable
attraction.
If the cone and the traverse are driven at constant speeds, the yarn speed required
to lay the yarn on the small-diameter end (B in Fig. 9.5) is different from that needed
240
Handbook of yarn production
A
A
B
B
Fig. 9.5
Cone surface
at the base (A). To make such a system work, there can either be intermittent yarn
storage or the traverse rate can be made to vary across the traverse.
Ideally, the yarn cone should be driven by another cone, to avoid scuffing of the
surface, because two cones mesh together perfectly. However, some winders use a
wide cylinder to drive the yarn cone which is kinematically incorrect. The cone tends
to run with its mean surface speed in synchronization with the drive cylinder; the
base runs faster and the tip runs slower, the differential movement causing scuffing.
Often, a narrow cylindrical rim is used to drive the cone. Although this avoids most
of the scuffing, it causes some local compression on the surface of the cone.
9.1.6 Traverse mechanisms
One common type of winder has a reciprocating traverse. As mentioned, by varying
the relative speeds of traverse and package, or their relative positions, one may obtain
a pattern breaking effect. A displacement of lay 1 with respect to lay 2 can produce
a pattern breaking effect as indicated in Fig. 9.2(c).
Care has to be taken to keep the final guide close to the surface of the package
otherwise the yarn lag can be troublesome and hard shoulders will be produced. The
yarn lag is due to the uncontrolled length AB in Fig. 9.6. When the guide is moving
leftwards, the yarn lies at an angle such as is shown at EF, but it lies at an angle such
as GH when traveling in the other direction. The lag tends to concentrate yarns at the
reversal points and give hard shoulders.
The simplest and a widely used form of yarn traverse is the grooved roller (Fig.
9.7). The grooved roller not only drives the package but it also lays the yarn on the
surface at approximately constant spacing. However, since the coil spacing is fixed,
there is ribboning every time the package reaches a multiple of the effective diameter
A
FG
B
H
E
Yarn at time 1
Yarn at time 2
Fig. 9.6
Yarn lag in winding
Post-spinning processes
241
Yarn package
or cheese
Yarn
Grooved roll
Fig. 9.7
Grooved roll traverse
of the grooved roll. There is also a fractional relationship as already explained. One
solution to the problem is to use a large diameter grooved roll, and another is to use
a pattern breaker. Some pattern breakers oscillate the package or the grooved roll,
sideways, randomly.
Local linear speeds at which yarn is laid onto the surface of a package, which
rotates at constant speed, vary with the angle of lay. If the surface speed is Vs and the
yarn speed is Vy, then at special positions, Vs = Vy, but elsewhere this is not so because
the yarn lies at an angle. One way to overcome the problem is to place a bow piece
in the plane of the yarn offtake (Fig. 9.8). The length of yarn between the supply and
the laying-on points varies in such a way as to take up the slack caused by the
variations in wind-on position and angle. In building a cone, more yarn has to be laid
on the base of the cone than on the tip. The angle of lay has to be biased away from
Yarn reciprocates over the bow
Yarn
Bow
Cheese
Fig. 9.8
The use of a bow to control yarn tensions
242
Handbook of yarn production
the sinusoidal and length compensation has to be introduced to prevent periodic high
tensions or overfeeds occurring. A special grooved roll or reciprocating traverse may
also be used.
9.1.7 Precision winding
The machines discussed so far have had a constant traverse rate and such machines
produce ribbons of denser structure at periodic intervals as the package grows in
diameter. There is a kinematic solution in which the traverse rate is varied. If we
differentiate Equation [9.1] with respect to time, we may restate the result as:
dφ/dt = K(πDU – mλV )
[9.2]
where dφ/dt = the rate of phase change of yarn layers, K is a constant, U = rotational
speed of the cheese in rev/s, V = the speed of the traverse in oscillations/s, and the
remaining symbols are as in Equation (9.1). If mV and DU are kept in synchronism,
dφ/dt = k, where k is another constant. If k is chosen appropriately, the ribbons can be
eliminated and a denser package can be made. Separate drives are needed for the
package and traverse and this makes the apparatus expensive.
Some machines use a precision wind in which the diameter is sensed and the
traverse speed is adjusted to give a constant value of the coil advance.
9.1.8 Direct winding
Where the yarn is wound directly on the cones or cheeses at the yarn manufacturing
stage (for example, in rotor spinning), great care is needed in setting the bow piece
and traverse mechanism. Unlike the case where the yarn is being withdrawn at will
from another package, the supply velocity is fixed by the process. As was seen
earlier, the velocity at which the yarn should be laid on the surface of the rotating
yarn package is not constant. If one requires a well-built cone with a considerable
cone angle, it is necessary to have a more sophisticated yarn storage and tension
control. An improperly set bow, or the operation of a bow piece system beyond its
limits, gives a poor package structure and a high frequency of end-breaks during
operation.
9.1.9 Winding tensions
Since yarns are visco-elastic in nature, any tension applied to them alters their
characteristics. Yarns need to be wound at controlled tensions because over-tension
damages the yarn; under-tension gives an unstable package of low density. Damaged
yarn performs badly as can be seen from the example in Fig. 9.9. Progressively
increasing the winding tension increases the tenacity at first and then it drops again.
At first sight, the increase in tenacity would seem to be an advantage. However,
careful examination will show that there is a progressive decrease in the breaking
elongation, which implies no improvement in the energy to break. Also, the weak
spots are strained more than normal, which is a bad feature. Another example, a
cotton OE yarn, showed that application of stresses up to 20% of the breaking value
increased the tenacity of the yarn between 1.0 and 1.5 gf/tex, but at the expense of the
breaking elongation. In weaving, any lack of elongation gives problems with warp
Post-spinning processes
243
Elongation
Tenacity
10
Tenacity
Elongation (%)
Tenacity (g/tex)
10
Elongation
5
5
0
Fig. 9.9
1
Winding tensions (g/tex)
2
Effects on yarn performance of high winding tensions
breaks.1 The value of such strength increases is therefore dubious and, without doubt,
over-stressing a yarn damages it. The effects on ring yarns are less because of the
more organized structure. Fortunately, OE yarns are rarely rewound.
9.1.10 Unwinding
The amount of yarn on a ring bobbin is insufficient for commercial use and many
ring bobbins contain faulty yarn, which has to be removed before final packaging.
The yarn is transferred to cheeses or cones and this involves the unwinding of the
ring bobbins as part of the transfer.
A frequent wind found among the supply packages is the filling or weft wind
(which typifies the ring package as sketched in Fig. 7.4(b)). The wind-on tension
varies throughout the wind because of variations in (a) tension during spinning, and
(b) the balloon during unwinding. High winding tensions are frequently met when a
ring bobbin is just started because of the smallness of the winding diameter, and this
is frequently a time when one sees an increase in the fault rate. Thus, a cheese or cone
has variations in yarn structure. Tensions change as yarn is taken from the beginning
or end of the ring tube; this changes the package structure. The appearance of thin
and thick spots further complicates the picture. It is not uncommon to find a thick
spot followed by a length of thin, over-twisted yarn. Even after piecings are cut out,
one finds that knots or splices are followed by thin, weak yarn with a different
appearance. There are a number of differences that are likely to show up in the final
product; often the problems are made worse by high winding tensions.
Bobbins from the ring frame have to be unwound at very high speeds as yarn is
transferred from the ring bobbins to the cones or cheeses. In the ring spinning process,
bobbins had rotated at high speed and new layers of yarn had been laid on to a plushlike surface created by the hairs being held out by centrifugal force. When the yarn
is removed during the winding process, the yarn tends to twist as it is removed; this
action removes fibers from adjacent coils on the bobbin. The consequence is a microunevenness that is not easily detectable on a spectrograph but is objectionable when
the yarn is assembled into fabric. The problem becomes worse as the ring frame
speeds increase. Also, it is a good reason to control yarn hairiness in spinning. Yarn
appearance is among the largest categories of customer complaint; it has been seen
1 Energy to break, i.e. tenacity × elongation, is the best criterion.
244
Handbook of yarn production
to rise at times as high as 40%. Even if a yarn has adequate strength, evenness, and
the correct twist, it is not always satisfactory. Rust and Peykamian [1] have shown
that the hairiness of yarns increases due to the winding process. During the winding
process, fibers are transferred from yarns on a supply package to the yarn just being
removed. Large transfers of fiber yield yarn faults; the normal transfer rate is less
than 10 fibers/m. The presence of yarn defects adversely affects unwinding performance;
it may also reduce winding efficiency, not only because of the increased clearing
needed but also because of the end-breakages in the process.
High speed unwinding of a package, whether it be a ring bobbin, cheese or cone,
gives a balloon that is chaotic, and peak yarn tensions can be very high. The difficulties
arise because the radius at which yarn is taken from the package varies rapidly
because of the build. Without any form of balloon control, these peak yarn tensions
severely limit the speed at which the yarn can be removed; this, in turn, limits the
winding speed of the whole winder. Balloon breakers are used to reduce the tension
variations and one of the latest types is shown in Fig. 9.10 [2]. All balloon breakers
increase the yarn hairiness and the mean winding tension. The take-off point of yarn
from the bobbin moves up and down the chase as yarn is pulled from the package;
this causes variation in yarn tension. The balloon breaker, which might also traverse
up and down, limits the tensions.
The yarn winding tension as it goes on to the new package is partly determined by
the unwinding tension from the supply package. It is further determined by the
additions and/or multiplication of tension (see next page) produced by the tension
controllers and guides. Some tension increase is inherent in the design of the system;
some is used as a means of control. The latter is dealt with later; for the moment we
shall concentrate on the involuntary portion. A simple balloon breaker can double the
Yarn
Before winding
Control tube
@ 1500 m/min
Conventional
@ 1000 m/min
Yarn balloon
0
Height, arbitrary units
Balloon
control
tube
Bobbin
100
200
Hairiness, arbitrary units
Fig. 9.10
Balloon control and yarn hairiness
Post-spinning processes
245
hairiness of the original yarn; it also increases the winding end-breakages unless the
winding speed is limited. Improved designs, such as those with a control tube, as
shown in Fig. 9.10, limit the hairiness increase [2].
Turning now to the voluntary tension controllers, there are two basic forms of
these, namely, addition and multiplication types. The former gives an output tension
equal to the input tension plus a frictional component; the latter gives a result, which
follows Amonton’s Law. Many tension systems incorporate elements of both and, in
general:
Tout = K1 µ F + Tin (1 + K2 eµθ )
[9.3]
where K1 and K2 are constants, µ is the coefficient of friction, F is the normal force
acting on the element in the tensioner, Tin is the input tension, Tout is the output
tension, θ is the angle of wrap in radians, and e = 2.718. Amonton’s Law states Tout
= Tin (1 + K2 eµθ ).
9.2
Yarn joining
9.2.1 Defect removal
Clearing of yarn to remove faults is a crucial role for most staple yarns. The market
rarely accepts yarn with numerous and large defects; to maintain price and reputation,
it is essential to remove all unacceptable slubs, thick spots, thin spots, etc. The level
at which they become unacceptable is a matter of agreement between the buyer and
seller of the yarn.
Yarn is sometimes wound at high tension so that weak spots are found during
winding rather than at some later stage. An automatic knotter or splicer is used to join
the ends created by the removal of the defect. To remove weak spots consistently, the
winding tension has to be uniform. In many cases it is anything but uniform and,
under such circumstances, some weak spots are missed; the efficiency of clearing is
thus reduced. Furthermore, some weak spots are overstrained and this might cause
problems later. Thick spots are often detected by passing the yarn through nub plates
that contain slots that will pass normal yarn but not the thick spots. The rise in tension
when a thick spot is caught in the nub plate causes the yarn to break; the ends are then
knotted or spliced to give continuity. Again there is a periodic over-stressing of the
yarn. Defects, and the consequential damage, are often concentrated in yarn taken
from the base of the ring tubes (because end-breaks in spinning are more frequent
there).
A satisfactory alternative is to use a sensor to detect the defects, and to use the
signals from it to actuate a cutter and a knotter or splicer. Optical or capacitive
sensors are frequently used. Accurate defect removal requires measurement of the
linear density (or ‘yarn diameter’) of the running yarn. When a bad spot which is
outside the prescribed limits passes through, the winding head should stop and the
section of bad yarn should be removed. Any newly cut or broken ends are then joined
before the winder recommences winding. A patrolling piecer assembly finds the ends
on the supply and uptake packages, splices the ends and then restarts the winding
head. A machine might contain tens of such piecers serving, perhaps, a hundred
winding heads. The procedure is automatic and needs no human intervention unless
(a) the winding head is improperly set, (b) the section of yarn is particularly bad, or
246
Handbook of yarn production
(c) there is a mechanical failure. The limits can be set independently to include the
thickest and thinnest allowable linear densities at various fault lengths.
Capacitive sensors are most commonly used to detect the defective portions of
yarns. Sometimes, however, optical devices are used, in which case one has to be
careful with the use of external lights around the machine unless the sensors work in
the infra red part of the spectrum. Optical sensors have difficulty in discriminating
between thick and fluffy spots (they are not sensitive to yarn mass). On the other
hand, capacitive sensors are moisture sensitive and they cannot discriminate between
changes in linear density, fiber composition, or moisture content. Both of the sensors
mentioned can perform satisfactorily providing that the moisture content is properly
controlled by appropriate air conditioning and that there is reasonable quality control
in spinning. Their use permits the setting of the winding tension to appropriate levels.
Supply packages should be conditioned for an adequate time because moisture
penetrates at a relatively slow rate; it is prudent to condition them for at least 24 hours
before winding. Since water is cheaper than yarn, it is necessary to define the moisture
content of the yarn in the contract between buyer and seller. It might also be pointed
out that many fibers swell when they imbibe water; this swelling and the subsequent
shrinkage can disturb the lay of the yarn on the package. Conditioning of finally
wound packages can lead to the production of hard shoulders. This leads to difficulty
in unwinding the packages and yarn damage. It is preferable to condition the yarn
before winding and then maintain the moisture content by appropriate control and
packaging of the final cones and cheeses.
The winder is used not only to join the component yarns as just discussed, but it
provides an opportunity to test the yarn for defects, cut them out, and join the new
ends. The consequential reduction in defects is of the greatest importance for quality
and for profitability of the mill. Clearly, if the clearers are set too close, too many
minor defects are removed and replaced by splices. There can be a deterioration in
quality because even the splices (as good as they may be) are still minor defects.
Equally clearly, failure to remove serious defects can have serious undesirable effects
on the trading relationships involved. The presence of unwanted defects may be the
subject of a complaint by the customer that may involve payment of a financial
settlement and a loss of confidence.
9.2.2 Splicing
Until the last decade, it was common practice to knot yarns together, but the knots
were a source of weakness and were defects in their own right. Nowadays, yarns are
spliced together using mechanical or air-jet splicers, which produce a joint that is
usually at least 70% of the strength, and generally less than 130% of the thickness of
the parent yarn. Splice efficiency is used as a measure of the strength of the spliced
portion of the yarn, expressed as a percentage of that of the parent yarn. The adoption
of splicing has greatly reduced problems in weaving, knitting, and dyeing.
There are two means of splicing in common use. One is in the form of a hand-held
device and the other is part of a winding machine. A single winding machine might
have, perhaps, ten splicers serving something in the order of 100 winding heads. A
typical winder/splicer makes between 10 and 30 pieces per package and ejects from
5 to 10 mg of fiber per package into the atmosphere. The splicers are complex,
expensive devices and there is a need to conserve capital by letting a single splicer
serve several winding heads. Also, that way, the splicer is kept in fuller employment
Post-spinning processes
247
than otherwise would be the case. Most devices involve air splicing but there are
some mechanical types. Both sorts can produce good splices providing they are
properly set. A review in the early 1990s of the industrial performance of a series of
splicers gave the results shown in Table 9.1 [3]. An old air splicer performed poorly
compared to more modern ones; this points to the improvement in piecer design over
a decade. Some spaces have figures in parentheses and these reflect the diminished
performance when a single machine (of the few needed) was poorly set. This is a
matter of personnel training rather than machine design. In a different survey of
several mills, the CVs of the settings of a given type of splicer were found to be
extraordinarily high. This is a bad sign as far as fixer training is concerned.
To make a satisfactory splice, the two yarn ends have first to be prepared to make
them properly tapered. Also, the fibers must be adequately separated and paralleled
so that they are capable of intermingling when the splice is made. Consider a typical
machine as shown in Fig. 9.11. Remember that the winders work at high speed, and
have to interrupt the winding to cut out the faulty portions of yarn before the new
ends are spliced together. It will be noted that scissors are provided to cut the unwanted
yarn ends after the two yarns have been laid in place. At this juncture, the ends of the
yarn are parallel and face opposite directions as shown in Fig. 9.11(a). Automatically
actuated clamps grasp the yarn at the appropriate places before the main splicing
Table 9.1
Performance of splicers
Splicer type
No of spinning
plants
% strength
efficiency
% bad
splices
% CV of splice
strength
Mechanical
Air (Maker 1)
Air (Maker 2)
Air (Maker 3)
Old air type
2
4
3
3
2
98
98 (94)
93
91 (81)
85
0
0 (8)
3
0 (27)
10
10
16 (19)
12
12 (21)
16
Yarn B
Yarn B
(a)
Ends being
spliced
(d)
Yarn B
(b)
(c)
Scissors
clamp
X
Spliced yarn
(e)
Air
Splicing
nozzle
X
Air
Splice
Clamp
Scissors
Yarn A
Time 1
Lay and cut
Yarn A
Time 2
Conditioning ends
X = End-conditioning nozzle
Fig. 9.11
Yarn A
Time 3
Form loops to
retract ends
Time 4
Splice ends
Stages in splicing
Time 5
Remove
spliced yarn
248
Handbook of yarn production
procedure begins. The free ends of the two yarns are sucked into end-conditioning
nozzles and air blasts are provided to condition them before joining. To condition the
ends, the yarns have to be gripped and fibers sucked from the exposed ends to taper
them (Fig. 9.11(b)).
A spectrum of fiber lengths is removed from the ends. The violent airflow causes
some fiber entanglement in the remaining fibers at the yarn ends, which makes it
easier for the fibers to intermingle during the splicing operation itself.2 Splicing is
carried out after the two conditioned yarn ends are laid inside the splicing chamber
so they are parallel, facing opposite directions and appropriately spaced without the
tips of the conditioned ends protruding. One way to do this is to withdraw loops of
yarn as shown in Fig. 9.11(c). The splicing chamber of an air splicer is sometimes
made in two parts which open to allow easy insertion of the yarn ends and then
permit closure for the splicing phase of the operation. The two ends are spliced
together by a rapidly rotating body of turbulent air inside the splicing chamber. The
turbulence is induced by air that enters the cylindrical chamber tangentially. The air
blast first intermingles the fibers and then causes the newly made joint to rotate to
produce false twist. The yarn is then removed from the splicer (Fig. 9.11(e)) and
winding is recommenced.
If properly restrained, the false twist at stage (d) accumulates on one side of the
splice until the air ceases to flow, at which time the false twist flows into the joint.
Twist distribution during the splicing operation is shown in Fig. 9.12(a), the yarn size
having been exaggerated to show the twist directions and level. This latter is important
because the twist in the splice gives the joint an appearance similar to that of the
parent yarn (and strengthens the joint too). Consider the case of a splicing chamber
designed for use with normal Z twist yarns. A temporarily high Z twist exists on one
side and an S twist (or very low Z twist) exists on the other while the air blast
operates. There is about 15 to 20 mm of zero twist between them then. After the air
blast stops, the twist is redistributed to give what appears to be reasonably uniform Z
twist throughout the splice. There is a distinct chance of damage to, or failure of, the
portion of the yarn with low twist during splicing. Also, if the tensions within the
zone are kept low, there is a probability of snarls forming in the high twist portion
during splicing.
Consider the nature of a splice. To avoid a thick splice it is necessary to taper the
ends to be spliced so that the joint is not obvious. These ends have to be in the proper
relative positions when the splice occurs. In the case shown in Fig. 9.12(b), the
tapered ends are misplaced to give a thin spot. This is also usually a weak spot, which
is undesirable on that account. The yarns could have been overlapped too much, in
which case there would be a thick spot and two undesirable splice-tails (Fig. 9.12(c)).
These tails catch up in knitting and weaving and are often the subject of customer
complaints. The splicer should be set to avoid these tails, even at the expense of a
slight loss in splice strength. Figure 9.13 shows two such bad splices, the top one of
which has a wrapper on the right-hand side and the bottom one a quite undesirable
tail on the left-hand side.
For these reasons, and to uphold quality standards, it is very important to maintain
the correct timings, settings, and tensions. The various motions described are performed
by a series of cams and levers, and it is essential that the components be well maintained.
2 If there are 50 splices per cone, the count is 20s (≈ 30 tex), and each splice converts about 1 inch
of yarn to fly, then 0.083 lb (≈ 38 g) of fiber is ejected per cone, which is a significant source
of pollution.
Post-spinning processes
249
Yarn 2
Yarn 1
Clamp
Clamp
Low twist
High twist
False twist
(a)
Splicing chamber
Thin spot
(b)
Thick spot
Unwanted tail
Unwanted tail
(c)
Fig. 9.12
Fig. 9.13
Splice structure
Imperfect splices of a dark yarn with a light one
Accumulations of lint, loss of lubricant, and wear on machine parts can alter the
performance of the machine and degrade the quality of the product.
The alternative method to air-jet splicing involves a mechanical device which
untwists the yarn end locally and pulls the tail away, causing the remaining fibers to
become parallelized. The distribution of fibers in the remaining tail gives a tapered
end as required. A shear field produced by two counter-moving disks is used to
produce the twisting action needed to splice the ends together. This mechanical false
twister has rubber elements, which apply pressure to the assembly during the splicing
operation and, again, careful maintenance is required to keep the system in good
working order.
250
Handbook of yarn production
Splices are commonly tested for strength, in the field, with a hand-held dynamometer.
The parent yarn on either side of the splice is normally tested also and so-called
splice ‘strengths’ (more properly, ‘splice efficiencies’) are usually expressed as (100%
× splice strength ÷ average of the parent yarn strengths). With a good splice, it is not
easy to distinguish the splice from the rest by eye. Nevertheless, if the yarn is looped
in the region of the splice, it will be seen that the splice is stiffer and does not
conform to the same radius as the rest. This is helpful in finding the exact position of
a splice that is about to be tested.
9.2.3 Winding machines
Most modern winding machines used in staple spinning incorporate not only the
function of winding but also that of removal of faults. They involve unwinding the
supply package (usually ring bobbins), splicing the yarns after cutting out faults, and
winding the package that goes to the customer or to some intermediate process. On
the other hand, winders for filament yarns are not concerned with the removal of
faults and splicing. All winding machines are designed to produce stable packages of
undamaged yarn at maximum speed.
A typical modern cone may well weigh more than 10 lb (≈ 4.5 kg) and can be
composed of yarn from up to 50 ring bobbins. The winding machines work at speeds
up to 1500 m/min (≈ 1370 yd/min) and the process can impose considerable strain on
the yarn if the machine is not properly set. The surfaces touched by the yarn are
usually coated or treated to reduce friction and abrasion. Increasingly, ceramic inserts
are being used at critical places to reduce the wear on the machine. The minimization
of the wear and tear on the yarn is also important; however, it is not always easy to
detect the damage until it enters a subsequent process. Wear can produce hairiness,
nep, and other appearance problems. Consequently, maintenance is essential.
Another function of the winder can be to meter the amount of yarn on a package.
This can be important to a customer, because it is desirable that all yarn packages in
a creel of a warper or knitting machine should run out at the same time. In this way,
there are few or no remnants of yarn left on the package centers to become waste.
Winding is so near to the customer interface, that it should be taken very seriously.
The author’s experience in the 1990s was that roughly 25% of claims made on
complaints arose from winding problems. This figure is not fixed; it is amenable to
change through improvements in technique and equipment.
9.3
Ply yarns
9.3.1 Plying
For sewing threads, as well as certain speciality and industrial yarns, it is necessary
to ply (i.e. to double or fold) the yarns to give them a smoother and less hairy
character. Doubling improves the evenness; plying balances torque if carried out
correctly and binds some of the hairs on the component yarns. The traditional methods
include assembly winding (Fig. 9.14) to place the single yarns parallel to each other
as a closely spaced pair (or group) of yarns on an intermediate package. The new
package is then used as a feed for a twisting machine and the output is a plied yarn.
However, the cost of assembly winding approaches 25% of the total winding costs
and the system is prone to problems. If one of the ends breaks in the process, then
Post-spinning processes
Fig. 9.14
251
Assembly winding
only a single yarn is wound until the machine is stopped. Another problem is that an
end from an adjacent package can be caught up to form a three-fold (or four-fold),
instead of the desired two-fold, yarn. Such faults, if not spotted and removed, are a
sure cause for complaint by the customer. The lengths of yarn on every package are
best matched to avoid wastage. Such practice also eliminates the need for the existence
of partly consumed feed packages on the winding frames and is desirable since the
presence of surplus packages increases the risk of the three- and four-fold yarns just
discussed. Because of these problems, it is modern practice to wind and clear the ring
yarn so that standard cones may be used for the feeds. This may not be the most
efficient way of converting the yarn, but the field is not large enough for there to be
promise of much machine development.
9.3.2 Singeing
Where smooth cotton yarns are required, as in sewing thread, they are sometimes
singed (or ‘gassed’) to remove the hairs. Such yarns are often two-ply; long-staple
cottons are frequently used to give maximum strength and resistance to abrasion. A
micrograph of a typical ring yarn before and after singeing is shown in Fig. 9.15. To
singe the yarn, it is passed through a flame at a steady speed; the rates of fuel gas and
air are carefully adjusted so that sufficient hair is removed without damaging the
core.
252
Handbook of yarn production
Fig. 9.15 Ring yarns before and after singeing
The process normally produces CO2 and soot. However, if the ratio of fuel-gas to
air is incorrect, the process will produce toxic carbon monoxide rather than CO2.
Therefore, care has to be taken to set the equipment properly.
Careful venting of the workspace is vital because cotton dust is highly flammable.
The singeing is carried out in a walled-off space otherwise there could be explosions
in the fly-laden atmospheres found in other fiber processing rooms. Other points to
consider: soot deposits can ruin otherwise perfectly good yarn packages, heat of
combustion needs to be removed, and the health of the workers must be preserved.
Singeing of yarns containing meltable fibers, or ones that char (like wool), is not
recommended.
9.3.3 Sewing threads
Although the manufacture of sewing threads does not necessarily involve new
technology, it does have special requirements for the yarn structure. Sewing threads
are often plied to give strength and uniformity to the product; they are often made
from cotton, but special threads of linen and silk are also produced. Yarn hairiness
has to be controlled to reduce the fly build-ups around the sewing needle area. To give
the desired strength and smoothness, cotton threads are often singed and then mercerized.
The latter is a process that swells the cotton fibers with caustic soda and stretches
them to improve molecular orientation. This, together with the high twists used,
makes them rather expensive. One of the reasons for using cotton is that it does not
melt. With the ever-increasing speeds of sewing machines, needle melts can be a
problem with many of the man-made yarns, especially with certain stitch constructions.
Thus, a source of high tensile threads is denied to some garment makers unless
special needle cooling arrangements are installed. There are some special man-made
fibers, such as Nomex, a heat resistant aramid fiber, but they are costly and this is not
a universal solution to the problem. Other solutions include the use of a man-made
filament core, sheathed with natural fibers to protect it and to give the desired
aesthetics.
In all cases, waxing of the yarns is essential. The wax must be sufficient to give
lubrication, even at the elevated local temperatures in the needle eye. However, the
quantity must not be so much as to cause clogging of the guides and needle eye.
Problems can also come from differences between the yarn and fabric in the
matters of elongation, tension differences, and dyeability. If the thread contracts more
Post-spinning processes
253
than the fabric being joined, the seams pucker. If the dye affinity of the thread differs
from the fabric, it becomes visible.
9.3.4 Other plied yarns
Plied yarns are used for purposes other than sewing threads. Plied yarns are not only
more even and of enhanced strength, they are durable, flexible, and low twists may be
used to give a soft hand. Acrylic and wool yarns are often plied for the hand knitting
market; this is because of the soft hand that can be created. Aramid yarns are plied
in various complex structures to give strength to ropes and load-bearing strands.
Thus, plying is widely used within certain specialty markets, but for run of the mill
products the process is too expensive. The improvements in quality of singles yarns
over the last half-century have undermined the broad plied yarn market.
9.4
Automation
9.4.1 Patrol theory
Consider the case of manual winding. To optimize the use of a winder, a relationship
has to be established between cost and the number of winder spindles to be assigned
to a worker. If the assignment is too low, the machine efficiency is high, but the
operator is not fully occupied and the operator efficiency is low. If the assignment is
too high, spindles stand idle waiting for the operator and the machine efficiency is
low. It is necessary to balance the two forms of efficiency so that the overall cost is
at a minimum. For simplicity, ignore the cost of doffing the large output package and
any waste. Assume that the operator progresses steadily in one direction, and only
repairs end-breaks or replaces empty input ring bobbins by full ones as he/she comes
to them (the interventions are called ‘events’). Spindles that have been passed have
to wait until the next circuit. The theory is similar to that which applies in spinning
(see Equation [12.1] in Chapter 12), except that the number of events/hr, E, replaces
the breaks/hr, B, and some terms are neglected. Suffice it to say that the estimated
optimum winding assignment, aw, is given approximately by:
aw = √{CLE/(VtCfw)}
where E
V
CL
Cfw
t
[9.4]
= number of events/yd
= velocity of yarn in yd/hr
= labor cost/hr
= Fixed cost/hr for winding machine
= average time in hrs to complete one patrol during which time all
spindles in the assignment are inspected once (including workbreaks).
The assignment is strongly affected by the pre-existing number of faults in the
yarn. The spindle speed Vw, the patrol time t, and the cost ratio CL /Cfw play a significant
part. The performance of the winding department is strongly affected by the preceding
operations. This estimate can be used where the winder transfers yarn from one large
bobbin to another but with ring yarns, manual winding is onerous and expensive. It
is normal to then use automatic winders.
254
Handbook of yarn production
9.4.2 Automatic winders
With automatic winders, the knotting or splicing function is taken over by a traveling
automatic device which searches for the broken ends, makes a join, and then sets the
spindle in production again. Operator attention is only required when the robot has
failed to make the join after several attempts; a signal light draws attention to such a
condition. The robot treats the exhaustion of a feed package in the same way as an
end-break, except that, in this case, a new feed package is presented automatically.
Various functions are carried out by separate systems that form part of the winding
machine. The traveling splicer operates independently of the other functions of the
winding machines. It patrols to and fro along a fixed path searching for end-breaks
or exhausted supply packages. When there is an exhausted ring tube, another mechanism
replenishes the magazine and then leaves it to the knotter or splicer to make the join.
There is also a mechanism that detects when the output package needs doffing,
substitutes an empty core for the full package, and restarts winding on the new core.
Consider the supply system. Several bobbins of yarn should be available to each
spindle. When there is a need for a fresh supply, the machine will discharge the empty
bobbin and bring a fresh bobbin into place. It will then find the end and join it to the
end from the delivery package before continuing to wind. It will do this without
assistance from the operator except when the machine malfunctions. This enables the
operator to distribute his or her work amongst several machines; it is only necessary
to ensure that the turret or magazine has a sufficient supply of bobbins. In modern
machines, this too has been automated. A hopper of ring bobbins supplies them to an
automatic device, which aligns them, rejects bad ones, and carries the aligned bobbins
to the individual spindle magazines.
Suction arms move to the surface of the supply and/or uptake packages to find the
ends lying on the surfaces of the packages when the operation has to be restarted. A
restart is needed when a new supply package is introduced, an output package is
doffed, or an end-break is detected. These ends are inhaled into the suction arms for
the limited duration of the transfer. During this time, yarn is consumed, but it is under
a controlled tension that makes the laying of the yarn into the piecer manageable. The
suction arms move to position the yarn as is shown in Fig. 9.11(a), so that the splicing
can continue. After the splice is made, the machine restarts automatically. If there is
a failure to splice, the machine attempts to splice again. However, if the number of retries is beyond the pre-set limit, a warning light appears to notify the operators that
the particular winding head requires attention. The frequency of appearance of warning
lights is often used as a crude measure of the quality of the yarn leaving the ring
spinning machine. The operator has a trouble-shooting function rather than that of a
server; the operator assignment is different from that for manual winding. Similar
factors operate in determining the assignments but the coefficients differ.
The capital costs of the knotter or splicer relative to the rest of the machine should,
in theory, determine the assignment of the robot. The mathematics of optimum operator
assignments applies (see Chapter 12); the theory indicates the importance of keeping
the yarn joining frequency within the economic limits of the machines. This implies
that the CV of the yarn and the defect levels should be carefully controlled so that
they do not exceed the design level (which is relatively low). The defect level of the
input yarn directly affects the number of joins per package, as well as the value of the
yarn. The number of joins should be at a minimum. Generally, a poor performance in
winding is a sign to expect poor spinning and future problems in fabric manufacture.
Many problems arise in the early processes, and clearing acts as a sort of filter to
Post-spinning processes
255
buffer the impact of these early problems, but does so at a price. Clearing does not
completely remove all effects of the faults produced in the early stages; the process
of clearing is carried out at the expense of winding efficiency. Whilst it might be
better to clear yarn and accept the low winder efficiencies than vice versa, it is better
still to spin good yarn in the first place.
9.4.3 Linking
Recent practice was for spinning bobbins to be gathered in tubs as they came from
the autodoffer. Tubs full of bobbins were placed randomly in the hopper of the
winding machine, which sorted the acceptable bobbins from the rest. The accepted
ones were then passed to the winding heads for the winding process. In the process,
the bobbins suffered considerable jostling. There is a trend nowadays to connect a
dedicated winder to a ring spinning machine. This has advantages and disadvantages.
The bobbins pass smoothly from the autodoffer to the winding head without being
subjected to jostling and possible damage. The transfer is automatic and saves labor
costs. It is also possible to keep track of the source of faulty bobbins, which is of
great value in quality control if appropriate records are kept and used. However, the
productivity of the ring frame is highly dependent on count and the winder is less so;
thus, there is a balance point between ring frame and winder at which the productivities
match. Departure from that balance gives an excess of winding heads or a lack of
them. The latter cannot be contemplated; the result is that a mill usually has excess
winding heads. Some machines are made so that the heads can be moved from one
linked system to another to ease this problem.
9.5
Two-for-one twisting
To make ply yarns, it is necessary to twist two or more singles yarns together.
Following the assembly winding stage, paired yarns are often twisted using two-forone twisters. (There are still ring twisters in use, but they are expensive to run.) The
principle is shown in Fig. 9.16(a). Some yarn makers use two-for-one twisting without
assembly winding, where two packages are mounted inside the balloon (Fig. 9.16(c)).
However, the process is not as efficient as assembly winding because more space is
used up inside the balloon envelope and the winding continues less smoothly. Figures
9.16(a) and (b) show two methods of controlling the yarn balloon, which surrounds
the yarn package(s). In one case, a circular balloon rail is used and in the other, a
cylindrical pot. This is analogous to the conventional balloon control ring in a ring
frame. It may be compared to the control pot shown earlier in Fig. 9.10. The yarn is
protected by smooth metal pots in order to facilitate start-up, to separate the yarns,
and to protect the feed yarn from the ever present fly. Plastics tend to become damaged
in use and stainless steel has been found to be satisfactory. To conserve valuable
space inside the balloon, it is preferable to use precision wound packages because the
package density can be raised by about 25% when compared with normal crosswound packages. However, this is not always feasible, especially if the cheaper rotor
spun yarn is used as a feed. Rotor spun yarn is frequently used because the packages
from the spinning machine need no winding before two-for-one twisting. Sometimes,
for very fine yarns, up-twisting (see Section 3.3.4) can be used instead, because the
256
Handbook of yarn production
Pigtail guide
Yarn passes down through the
spindle and then up outside
the balloon rail
Rotating flyer
Neither balloon rail
nor package rotates
Rotating whorl
Tape
(a)
Central tube
Yarn
Package
Pot
(Stationary)
Storage disc
Outer pot omitted for clarity
(b)
Fig. 9.16
T
S
Yarns shown grossly oversize
(c)
Several forms of two-for-one twisting
damage to the surface layers of the yarn is less. The increase in cost is a reasonably
small percentage of an already expensive yarn.
The package(s) inside the balloon in a two-for-one twister is/are usually held in
place with magnets, which act on armatures in the yarn package center. This enables
a vertical spindle arrangement to be used which conserves space. Any rotating member
in the magnetic field has to be non-conducting, otherwise there is an energy loss and
some local heating. Yarn tension is controlled by a choice of various spring-loaded
and ball-type devices, mounted in the central hollow shaft of the winder through
which the yarn travels. The use of various pots not only protects and separates the
yarns, but provides some protection against balloon collapse. Tensions in this type of
twisting do not vary greatly when everything is properly adjusted. All surfaces in
contact with the yarn need to be polished and to have an immunity to wear and ill use.
Post-spinning processes
257
It is normal to have a storage disk S under the rotating throw-off plate and a yarn
tensioner T as shown in Fig. 9.16(c). The angle of wrap on the storage surface plays
an important role in determining the tension. If the tension leaving the tensioner in
the hollow spindle becomes too low, the balloon becomes unstable. In that case, the
rate of wrapping the storage disk (Fig. 9.16(b) and (c)) increases, with the result that
the tension in the balloon rises due to the increased capstan effect. Conversely, too
high a tension has the opposite effect. The maximum yarn tension is at the balloon
node because the energy for yarn movement along its axis comes from the winder.
(This differs from the balloon in a ring machine where the energy for yarn translational
movement comes from the bottom of the balloon.) The winding portion of the machine
must be controlled (a) to preserve constancy of the twist level, (b) to be independent
of variations in tension arising from the twisting unit, and (c) to produce a package
structure suitable for the end use. In other words, it must be possible to unwind the
yarn at speed and give a product as nearly uniform as possible in all respects. This
requires a good tension controller and lay mechanism to produce the necessary structure.
Twisting machines are often used for waxing and this is usually done to meet the
needs of the knitter, although the wax can help in the winding process also. Passage
over guides and machine surfaces tears out fibers from the surface, especially when
frictional forces are high. Waxing and plying are both methods of limiting that increase
in hairiness and in wild fibers. Damage created in twisting also results in the generation
of fly and dust, which brings other quality control problems. The use of correct
speeds and judicious amounts of lubricant limit the problem.
9.6
Customer concerns
9.6.1 Yarn contraction
When a strand such as a yarn is twisted, it tries to become shorter. If the yarn were
allowed to move freely, no tension would be generated and the yarn would become
shorter due to the twist; this is known as ‘twist contraction’. A typical value for twist
contraction for a typical 30s cotton ring spun yarn is about 4%. Contraction of
twisted yarn after it is wound on a package generates tension, which causes the
package to become more tightly packed. Absorption of moisture causes contraction
in hydrophilic yarns because the fibers swell and need more room. Thus, changes in
moisture content of such packages cause variations in package density, and sometimes
make it difficult for the customer to unwind the package as was discussed in Section
9.2.1. The effect is greater with highly twisted yarns. Yarn is sometimes conditioned
by steaming to adjust the package weight to standard conditions if it is too dry to
meet the agreed specifications. Over-conditioning can make the yarns absorb too
much moisture. Not only does over-conditioning affect the weight of the yarn packages,
it affects the tightness of the package to the point where it creates the difficulties for
the customer in unwinding as already discussed.
9.6.2 Winding and yarn dyeing
Dyeing of yarn is often carried out in an autoclave similar to that described elsewhere.
Dye is forced through the package under pressure and the whole autoclave is held at
a sufficiently high pressure to gain the necessary dyeing temperature. An autoclave
is an expensive piece of equipment in which the yarn packages have to be closely
258
Handbook of yarn production
packed to fill the working volume of the steam chamber. (An autoclave is briefly
described in footnote 1 in Chapter 4) Usually a series of cheeses or cones is mounted
on a mandrel and several of these loaded mandrels are inserted into the autoclave,
with their axes parallel to one another. The mandrels are usually part of the autoclave
itself because the dye solution travels through the perforated hollow centers on its
way to or from the dye packages. Frequently the direction in which the dye is pumped
is alternated; inside to outside, outside to inside, and so on. As might be deduced, the
dye pressure depends on the steam pressure in the autoclave, as well as on the density
of the packages. Low density packages tend to have larger passageways through them
than more dense ones. Thus, the package structure affects the dye pressure needed.
The yarn package transport system for these products should be designed with the
capacity and arrangement of the autoclave in mind.
9.6.3 Preservation of yarn tails on packages
Cones and cheeses are usually made with yarn tails from the inner and outer layers
of the package secured at some agreed upon position on the package centers. The
purpose of this is to enable the head of one package to be tied to the tail of another.
Thus, when the first is exhausted, yarn is automatically taken from the next, and so
on. Packages of this sort are frequently used in creels of machines in which many
short chains of packages supply whatever equipment the customer is operating. Within
these chains, the packages are tied head to tail. This permits the labor of tying the
ends to be deferred by the operator because the packages actually supplying the
machine act as reservoirs; it makes possible the continuous operation of the yarn
using machine. It is commercially important that the tails always are present in the
form specified by the user. Winding machines need careful adjustment to produce the
desired tails reliably. Even if the packages are used in-house, the receiving department
should be treated as a customer.
9.6.4 Shipping
It might seem so obvious that it needs no statement, and yet mills lose thousands of
dollars by errors in shipping. In business, there is a time and efficiency factor in the
goodwill generated. If the goods do not arrive on time, no matter that they are
perfectly satisfactory and reasonably priced, the shipments might be returned. This
brings in no revenue to the spinner and only produces extra costs that have to be
subtracted from any profit. If the goods are shipped to the wrong place, or the wrong
items are shipped, the effects are equally disastrous. It is imperative that the shipping
operation be efficient and effective.
The relatively new idea of just-in-time (JIT) shipping is one where incremental
orders are transmitted electronically, from user to supplier, as required; the supplier
ships goods in timely quantities to keep the user’s operation going without large
inventories. Inventories would otherwise consume working capital and JIT reduces
interest charges and fees. However, there is a downside to the scheme. A sales system
has inertia; some of it is due to the inventory stored in various parts of the supply line.
JIT seeks to reduce this cushion of stock; but in times of sudden demand, the lack of
a cushion can make severe demands on the primary producer, both in the shipping
and the production departments. Rather than lose business, many primary producers
keep an inventory to meet the need, but it involves risk and cost. The cost tends to be
Post-spinning processes
259
passed down the line because, if one business fails for these reasons, competition is
reduced and prices tend to rise.
References
1.
2.
3.
Rust, J P and Peykamian, S. Yarn Hairiness and the Process of Winding, Text Res J, 62, 11,
685–9, 1992.
Muratec Advertisement, Int Text Bull, 1, 21, 1996.
Lord, P R. Unpublished data from the author’s private records
10
Staple systems and modified yarn structures
10.1
Yarns of complex structure
Traditional ring spun yarns consist of fiber spirals and parts of each fiber approximate
to helices of the same pitch. Yarns of this sort can be untwisted to give roughly
parallel fibers. The qualifiers inherent in the above definition arise because of lateral
fiber migration, fiber hooks, and convolutions created in processing. Despite such
distortions, a fully untwisted yarn possesses very low strength. Processes such as
open-end spinning, air-jet spinning, and other specialist systems produce a structure
that contains varying pitches of the helical fiber segments. As a result, they can never
be untwisted to a point where the fibers are all roughly parallel. The yarns thus
possess significant strength when untwisted to give a minimum value of strength.
The variety of structures is wide. Structures for rotor spun and air-jet yarns are
described in Chapter 7 and Appendix 10. The process of rotor spinning also is considered
in Chapter 7.
10.1.1 Composite yarns
A series of developments related to the various specialist systems have been made in
which staple yarns and filaments are combined to give composite yarns. Usually, the
filaments are used as the core of the yarns, and staple fibers make up the surrounding
sheath. In this way, the filament yarns are placed where their strength is of the
greatest advantage and the staple fibers are placed where they can have the greatest
aesthetic value. Unfortunately, filaments are relatively expensive and so increase the
cost/lb of the yarn. Also they are often shiny, so that if the sheath does not cover the
core properly, the shiny filaments ‘grin’ through the cover to give streaky effects in
the final fabric. A description of one technology to produce composites by wrapping
is given later is Section 10.6.
Staple systems and modified yarn structures 261
10.2 Processes using modified twist
10.2.1 Processes using false twist
The advantage of using false twist is that there is no need to rotate a yarn package to
put in the twist needed, but the twist created is transient. The following sections
sketch two examples of how difficulty can be circumvented. The first example is the
air-jet spinner, which encourages a structural change in the yarn before the false twist
is released (Section 10.4). The second example is of self-twist yarns in which one
yarn with an alternating twist is plied with another to make a stable ply yarn (Section
10.7).
10.2.2 Processes using another form of modified twist
Streams of untwisted fiber can be encouraged to merge in such a way as to produce
a structure that looks like a ply yarn (one might call it a mock ply). It does this by
exploiting the conservation of torque. Torque applied to the outgoing ply can be made
to transfer to the ingoing component strands to create a ply twist (Fig. 10.1). (Reactions
R occur at the front drafting rolls.) The result is similar to a regular plied yarn and it
is made without the mechanical complexity of the traditional plied yarn operation.
The yarn structure is either an S-on-S or Z-on-Z ply rather than the conventional
balanced torque type of S-on-Z or Z-on-S. The yarn and process is described in
Section 10.7.
10.3
Compact spinning
10.3.1 Fibers in the twist triangle in ring spinning
The twist triangle in ring spinning determines much of the character of a ring yarn.
As discussed in Section A5.2.1, each fiber leaving the nip of the front rolls of a
conventional drafting system has a tension that depends on its lateral position with
respect to the rolls. There a wide distribution of fiber tensions and fibers traversing
at least one of the two selvages of the triangle exist at a high tension. Fibers in the
central zone often go slack. Cameras using short duration exposures have produced
images which have shown so-called wild fibers emerging from the front roll nip of
the drafting system and these are also slack. Parts of these wild fibers exist in space
remote from the twist triangle itself and contribute to the hairiness of the yarn. Fibers
R1
T1
Strands
from
drafting
system
Tout
Yarn
T2
R2
Tout = T1 + T2 + Losses
T1 + T2 = R 1 + R 2
T = Torque, R = Reaction
Fig. 10.1
Torque balance in mock ply
262
Handbook of yarn production
under high tension migrate to the center of the yarn as they enter the twisted transition
structure at the output of the triangular zone. Slack fibers migrate to the outside of
the yarn and form part of the loose hairy surface. The structure of fibers within the
twist triangle changes continuously (although for simplicity, descriptive models often
assume that the average tension distributions are symmetrical about the center-line of
fiber flow but this is not always true). Consequently, there is good reason to try to
control the distribution and variability of the fiber tensions within the twist triangle.
10.3.2 Processes of fiber control
If the fiber tensions can be controlled externally by a system of restraints, it becomes
possible to reduce the migration of what have been slack fibers and which might
otherwise have formed a loose structure at the surface of the yarn. Such a constraint
would be expected to produce leaner yarns with well-organized surfaces, in which the
outer fibers can bear a larger percentage of the load applied to the yarns.
Consider a system containing a perforated surface with fibers flowing on one side
of the perforations and suction applied to the other side. Friction forces will be
generated between the fibers and the perforated surface, which increase the fiber
tensions by a small amount. If the flowing fibers are those in the twist triangle, and
sufficient suction is applied, there should be few, if any, slack fibers entering the twist
transition zone and the tendency for what would have been slack fibers to migrate to
the surface will be lessened. Furthermore, if the suction apertures are controlled to fit
the shape of the twist triangle, the number of wild fibers can be controlled, which
again tends to reduce the loose structures mentioned and to reduce hairiness.
There are several possible arrangements that could fit this specification. A few of
them are: (a) a perforated hollow front roll with suction applied to the inside, (b) a
‘back-to-front’ perforated apron projecting from the front roll under (or over) the
twist triangle with suction acting through the apron thickness, (c) an option similar
to (b) but with a slot or aperture in a cover plate to control the position of application
of the suction. Several such designs were shown in the 1999 ITMA machinery exhibition.
Possible drawbacks to such designs are:
1
2
3
4
When an end breaks, it is more difficult to apply the conventional pneumafil
devices because the two suctions compete with the result that the dangers of a
lap-up are greatly increased.
Some machines have to be specially engineered, which increases the cost.
Accumulations of fiber debris from fibers entrapped in the perforation are likely
to increase maintenance problems.
Ineffectiveness for heavy yarn counts.
Relating to (1) above, the use of a roving-stop system to solve the lap-up problem
increases the cost of the machine, but the roving-stop mechanism does have the
advantage of eliminating pneumafil waste and all of the problems involved in recycling
it. Advantages include the possibility of producing smooth strong yarns, which would
be of interest to weavers because of the reduction in hairiness, coupled with a modest
increase in strength. This would make beaming, slashing, and weaving easier and
more efficient. Another area that could benefit would be in thread production, where
it might be possible to dispense with singeing. This suggests that, if such devices
become commercially viable, the first users are likely to be producers of yarn for
weavers or thread makers.
Staple systems and modified yarn structures 263
10.4
Air-jet spinning
10.4.1 The principle
Air-vortex and air-jet developments led to air-jet machines, which are not truly openend spinning machines but are related. In OE spinning there is an open-end, which
can be rotated, whereas in some of the yarns about to be discussed, continuity in flow
is given by a core. Fibers outside that core can be rearranged and trapped in the
structure to give different yarn characteristics. Götzfried [1], and later Pacholski [2],
showed that air-jets entering tangentially with respect to the bore of a nozzle cause a
vortex within it, and the high speed rotation of the air can be used to twist yarn
passing coaxially through the vortex. The pure air-vortex spinners did not succeed
commercially but they laid the groundwork for the modern air-jet spinning system.
They also laid the groundwork for some of the textured and composite yarns. If the
jet in the nozzle is inclined in the direction of yarn flow, it can help transport the yarn.
This is by virtue of the extra increment of yarn tension generated due to the axial
component of the air drag between the air and the yarn. The rotational speeds of the
vortex can approach a million r/min but the yarn rotation is likely to be limited to a
region around 200 000 r/min [3]; there is a large potential for high speed yarn
production.
An important development was that of the ‘fasciated’ (wrapped) yarn principle
[4]. The original idea was based on the addition of fibers to a flowing, false twisted
structure followed by the removal of torque at the exit of the false twister, which
causes the added fiber to be reverse twisted as shown in Fig. 10.2. Closely related to
this was the idea that the hairs on an incompletely formed yarn could be wrapped
about its core. Such hairs, entrapped in the structure, give enhanced cohesion to the
strand, even after untwisting. In patents by various authors [5–8], several sets of
apparatus and processes were disclosed which produced similar effects. The most
important of these will now be discussed.
In some examples, which are attributable to Murata [5, 6, 8], two twisting devices
were used, one to produce, say, S twist at the exit of the twist triangle, followed by a
device producing the opposite twist. The second device removes the false twist from
False twisted yarn
False twist is removed as
the yarn leaves the twister
Added fiber
Fiber locked on the yarn structure before untwisting
Resulting fasciated yarn
Fig. 10.2
Fasciated yarn
264
Handbook of yarn production
the core and completes the wrapping of the outer fibers about the core. The twisting
devices shown are air-jets, but other devices can be used. Directions of twist can be
reversed to produce a yarn that simulates the opposite yarn twist but this requires
different nozzles or settings. The common use of such machines is for the longerstaple fibers in the short-staple range. For example 1.5 inch (38 mm) polyester or
cotton fibers spin well.
It is normal for the wrapper fibers to be in the Z direction. In air-jet spinning, the
idea is to use the twist triangle leaving a roller drafting system to produce a very hairy
intermediate. The portion of the fiber flow that makes the core of the yarn, and the
extent of the yarn hair, are unlike the corresponding values with conventional yarn.
This is because the feed ribbons are much wider and the fibers in the triangular zones
(marked H in Fig. 10.3) are a greater proportion of the total than normal, and there
is a greater degree of drafting there. Zones marked T contain fibers under tension
derived from the pull of the air-jet. The zone marked M contains fibers that go slack
due to their shorter path length (when compared with the others) between the nip of
the drafting rolls and the vertex of the triangle. Slack fibers migrate laterally in the
core structure to interlock it, give it fiber cohesion, and create fiber loops and hairs
(see also Appendix 5). Compared with a conventional yarn, there is a wider range of
fiber tensions distributed over a greater width, and this promotes migration and
hairiness. Hairs on the outside of the core at A in Fig. 10.3 are more or less autonomous
and make only a loose and easily disturbed sheath structure. Figure 10.4 shows fibers
emerging from a roll nip that have then been twisted by an air-jet; the fiber flow
coalesces into yarn (shown at Y) due to the twist and the yarn is taken off in the direction
of the black arrows. The separation of the surfaces of the rolls creates a depression along
Drafting system
Twister 2
Z twist
Yarn
Sliver
Air jet 1
S twist
Twist triangle
H
Hairy yarn
A
T
X
Tension
M
Motion
T
H
Torque
H = Selvage zone
T = Fiber tension zone
M = Fiber migration zone
Enlargement of twist triangle zone
Fig. 10.3
Air-jet spinning
Staple systems and modified yarn structures 265
X
Air
Air
Roll nip
Y
Y
Air
Air
Elevation
Fig. 10.4
View in direction X
Air and fiber flow in a roll nip
the nip zone that induces airflow (shown with gray arrows) into the nip and this
airflow aggravates the hairy condition. The diagram shows two views of this airflow,
and the limited view in direction X also shows how disorderly the fiber mass can be
in this region. The major part of the twist triangle cannot be seen because it is masked
by the top roll. Similar, but much clearer, pictures were obtained by Jones [9].
An extension of the principle just explained is to feed two adjacent yarns being
made on an air-jet machine to a single take-up mechanism to create an assembly
package (Section 9.3.1). The yarn can be twisted later to make a ply yarn. This idea
has been used successfully on long-staple yarns and an example of it is given by the
Suessen Plyfil system, which can spin up to about 8 inch (220 mm) wool (and similar
fibers) from sliver in the range from 160 to 380 yd/min (150–350 m/min).
10.4.2 Machine design aspects of air-jet spinning
As mentioned, the hairs are important because they are laid on the core of false
twisted yarn leaving the twist triangle; the false twist is removed with the hairs in
place. The spinning action wraps the hairs around the core and there is enough lateral
fiber migration to lock the structure. The final product has little or no twist in the
core, but has a twisted sheath, which gives the structure integrity. Leaving aside the
single nozzle versions of air-jet machines, the false twist and rearrangement of the
sheath fibers are usually carried out by two air-jet nozzles set in line, close to the
drafting system. However, it is possible to replace the second nozzle by a mechanical
twister. The entry of the air-jet orifices has one component angle tangential to the
cylindrical main channel through which the yarn moves, and another component
angled relative to the axis of yarn flow. The latter is an important parameter because
it helps define the relationship between the twisting and linear translation speeds.
Oxenham and Basu [10] showed that if the jet orifice was inclined more than 60° to
the axis of the yarn there was difficulty in spinning, but at 45° spinning went on well.
The diameter of the vena contracta of this channel was usually 1.6 mm and the orifice
was 0.5 mm. In their experiments, the frictional characteristics of the chambers of a
first nozzle were altered by coating the surfaces with PTFE to give a low coefficient
of friction. Another nozzle was made from a ceramic material to provide a higher
coefficient. The PTFE coated nozzle produced the best yarn tenacity and the ceramic
266
Handbook of yarn production
one produced the greatest CV of tenacity. Air pressures of up to 3 kg/cm2 were used
in the first nozzle and 4 kg/cm2 in the second.
Air-jet spinning machines with more sophisticated drafting, fiber reconsolidation,
and twisting systems than those shown have now become established. Table 10.1
shows some data relating to different fiber finishes and it will be seen that finishes
with high coefficients of friction give poor results.
New machines with mechanical twisters following the first fiber consolidation
nozzle have appeared. The mechanical twister, which replaces the second air-jet false
twister, is formed by pairs of rolls with their axes crossed (Fig. 10.5). This layout of
rolls gives torsional as well as transport components of motion to the yarn as it leaves
the consolidation nozzle. The greater force acting on the hairs to press them towards
the center of the yarn increases the chance of integration into the yarn structure. One
certain result is that this arrangement produces less hairy yarn and possibly increases
the range of fiber length that can be spun successfully.
Air-jet machines often feature a five-roll drafting system with two pairs of double
aprons capable of drafts up to 400. The drafted strands are entangled or fasciated by
air-jets, as just described. The use of a wide ribbon passing through the drafting zone
helps enormously in this respect. In this way, a sliver-to-yarn system is possible,
which avoids the use of a relatively expensive intermediate roving. The machines can
produce cheeses or cones and thus separate winding costs are eliminated. Traditional
assembly winding is also avoided (assembly winding is discussed in Section 9.3.1).
Adjacent pairs of air-jet yarns can be laid side by side before winding to create an
assembly wound package and then pairs of yarn can be twisted subsequently, usually
by a two-for-one twister, to make plied yarns. Although variants exist, the basic idea
Table 10.1
Characteristics of air-jet yarn
Fiber length
(mm)
Fiber fineness
(d tex)
Yarn tenacity
(cN/tex)
Imp
per km
CV
(%)
Stops
per hr
38
38
38 a
38 b
32
1.7
2.2
1.4
1.4
1.7
18.0
15.3
15.5
15.0
19.5
280
–
48
20
40
15
16
14
14
14
18
38
85
20
40
Notes a = high friction fiber finish, Imp = Imperfection, b = low friction fiber finish, CV = CV of linear density.
Fig. 10.5
Crossed-roll twister
Staple systems and modified yarn structures 267
is to avoid the roving and winding operations and to produce the equivalent of
traditional plied or singles yarns.
10.4.3 Air-jet performance
Figure 10.6 shows a micrograph of an air-jet yarn made from 50/50 polyester/cotton
staple fibers, the polyester fibers having a length of 1.5 inches (≈ 38 mm) and the
cotton an upper half mean length of 1.05 inches (≈ 27mm). The wrappers that give
the structure cohesion are denoted by W in the micrographs. Polyester dominates the
wrappers and cotton is more prevalent in the core. This illustrates the importance of
fiber length. Long fibers produce longer hairs approaching the final twister and they
have a greater chance of becoming entangled with the core. Thus they generate
tension within themselves as they wrap around the structure in helical form. The
greater these tensions, the more radial forces are produced and the more cohesion
within the structure is produced. Short cotton fibers give problems with this sort of
spinning and long ones are relatively expensive. Nevertheless, a good reason for
persisting with cotton fibers is that consumers seek cotton yarns. A successful air-jet
process to make 100% cotton yarns is a worthwhile target. A 100% cotton yarn is
hairy and this suggests a deficiency in integrating the hairs entering the twister.
However, in general, it is claimed that the yarn defect level is lower than with comparable
ring yarns. Looney [11] showed that the fault rate for 100% polyester was reduced to
about 40% of that of the blend. He concluded that the result was, in part, due to the
relatively short, high micronaire cotton fiber used (0.93 inch (24 mm) and 4.24
micronaire). It is interesting to note that some of the traditional remedies intended to
improve quality sometimes do not produce the intended result. For example, increasing
the number of drawings from two to three caused a decrease of about 20% in the fault
rate as expected but it decreased the sliver cohesion by about 75%, which made the
handling of the sliver difficult. Mishandled sliver caused stops and faults. Intimate
blending can produce a reduction of up to 20% in yarn fault rates over drawframe
blending. Delivery rates up to nearly ten times that of a ring frame were reported.
Some idea of the performance traits of an air-jet machine when spinning polyester
yarns is shown in Fig. 10.7, based on data given by Looney [12] in 1984. The point
being made was that the yarn CV increases with the linear density of the fiber and the
yarn tenacity decreases. The system seems best suited for fine fibers.
In contrast, the work of Oxenham and Basu [10] showed that, when spinning
32 mm (1.26 inch) cotton of 26 g/tex fiber strength, yarn tenacities did not exceed
5 cN/tex (5.1 g/tex) and the yarn elongations at break were less than 8%. Thus,
despite the use of long cottons, there can be some problems in running 100% cotton,
although polyester/cotton blends are usually satisfactory.
W
Fig. 10.6 Micrograph of an air-jet yarn
W
Handbook of yarn production
Tenacity
Yarn CV
20
Yarn CV (%)
18
22
12 tex
X
20
20 tex
16
12 tex
X
18
Tenacity (cN/tex)
268
20 tex
14
16
10
Fig. 10.7
10.5
15
Fiber fineness (d/tex)
Air-jet spinning performance, 1984
Sirospun yarns and process
10.5.1 Sirospun yarns
As mentioned previously, worsted warp yarns are often doubled (i.e. plied). Plying
has a number of benefits: (a) plied yarns have a better evenness because of doubling,
(b) they weave more easily and this reduces costs in weaving [13], and (c) fabric
made from them is more durable and less likely to pill. Surface fibers on a singles
worsted yarn are sometimes loosely wrapped around the body; plying avoids difficulties
due to such wild fibers [13]. The Sirospun process makes S-on-S or Z-on-Z ply yarns
with somewhat similar characteristics to a normal S-on-Z ply yarn except that the
unidirectional structures can never be completely balanced.
10.5.2 The Sirospun process
A structure that is similar to the plied yarn can be produced on an ingeniously
adapted ring frame (Fig. 10.8). The yarn is called Sirospun. Two rovings (A and B)
are fed to a ring frame, with separators to ensure that each roving is drafted individually.
The two strands emerging from the drafting system converge into a single yarn, at J,
before they reach the lappet guide (otherwise known as a pigtail guide). The variations
in linear density and tension cause a random fluctuating twist to be generated in the
emerging yarn structure which gives it some of the characteristics of a plied yarn. If
a Sirospun yarn is untwisted, the individual strands have a low fluctuating twist,
which defines the strands such that, when they are twisted, they give the character of
a ply. Quite low levels of random twist generate enough surface fiber trapping to
improve weavability significantly; it is not necessary for the random twist to be
unidirectional for this purpose. A typical strand twist is of the order of 1 turn/inch
and this is sufficient to bind the wild fibers in place. The low twists result in potentially
high productivities. However, the system is restricted to long-staple fibers. A
disadvantage of the system is that if one strand breaks, long lengths of singles yarn
will be interspersed with the quasi-plied yarn. Such faults are known as ‘spinners
singles’. The difficulty is overcome by having a break-out device (shown at B in Fig.
10.8), which stops production when either of the two component strands is missing.
A suction then removes the fiber from the emerging strands. There is a cost to this
Staple systems and modified yarn structures 269
A
B
J
B
X
Fig. 10.8
Sirospinning
solution inasmuch as each spinning position has to have an extra piece of equipment
compared with regular spinning. Remembering that the productivity per spinning
position is still low, any extra cost of this sort is not trivial. On the other hand, the cost
of the detector is considerably less than that of a traditional plying operation set-up
on a comparable basis. It has the advantage that existing frames can easily be modified.
270
Handbook of yarn production
According to Lorenz [14], the Sirospun system has gained a respectable share of
the worsted market and it remains to be seen to what extent it can penetrate other
sectors. In short-staple spinning, the existing alternatives to ring spinning offer much
higher productivity and lower costs and penetration of that market will be difficult.
Lamb and Junghani [15] compared wool yarns made by the Sirospun system with
similar ones and found that the index of irregularity of the Sirospun yarns was
between those of conventional two-fold and Plyfil (air-jet) yarns. Three twist multiples
were tested (α = 73, 96, and 122) with similar results. The two-fold yarns were ring
twisted. They noted an increase in short-term unevenness in Plyfil yarns, which they
attributed to the high draft. The question of whether the twist levels could be reduced
was not resolved and it was pointed out that it is not the tenacity of the yarn that is
important but rather the weavability. Both the Plyfil and Sirospun yarns were torque
unbalanced, which was a disadvantage for knitters.
10.6
Hollow spindle spinning
10.6.1 Wrap spun yarns
The basic idea of a wrap spun yarn is for the machine to insert little or no twist in the
core and, at the same time, wrap another yarn or filament around the core at high
speed to make a composite yarn. The wrapper yarn or filament provides the forces to
compact the yarn structure. Frequently the wrapper yarn is a filament, which being
strong, can be wrapped at high speed without suffering the number of end-breaks that
would be encountered by a staple yarn.
10.6.2 Wrap spinning by the hollow spindle process
This technology provides a means of wrapping filaments about core yarns to enhance
the performance of the composite. Figure 10.9 is a diagram of a hollow spindle
system in which filament is taken from a bobbin mounted coaxially with the yarn Y.
A hollow spindle with a hook rotates about the same axis. The hook engages the yarn
and creates false twist above the hook, but the staple strand below the hook should
have little or no twist. The filament yarn, F, passes through the hollow spindle and
should have sufficient twist induced above the hook for the filament and staple
components to be brought into firm contact. The hook acts like an untwister similar
to a false twist spindle in texturing. Most of the false twist in the staple component
is removed as it passes through the hook. The twisting action causes the filament
to follow the surface of the staple component and the filament becomes tightly
wrapped about the very low twist staple core that emerges from the hook (yarn Y′).
Because of the high tenacity of the filaments, high production speeds are possible (up
35 000 r/min, which is about twice that of ring spinning). Sometimes, sliver-to-yarn
systems are used and the filament is wrapped around the drafted sliver. Occasionally,
both a filament and a staple strand (roving or sliver) are passed through the drafting
system to produce a bouclé or other effect. The system can handle short or long staple
but it is predominantly used for long-staple wrapped yarns. Xie et al. [16] created a
theoretical model and tested 64s wool yarns to find that yarn tenacities of up to 12 g/
tex were possible, with wrapper twists in the range 3 to 5 wraps/cm (≈ 1.2 to 2 tpi).
Staple systems and modified yarn structures 271
Staple strand
Drafting
system
Y
False twisted yarn
Filament F
Rotating hollow
spindle
Bearing
Y′
Hook
Wrapped yarn
Guide
Fig. 10.9
10.7
Hollow spindle spinning
Self-twist spinning
10.7.1 Self-twist principle
If a pair of worsted rovings are drafted and the emerging strands are passed through
a twisting system such as is shown in Fig. 10.10, a plied yarn is produced in which
both the ply and strand twists alternate. The emerging strands are called self-twist
(ST) yarns. The torque of the freshly emerging strands from the front rollers of the
drafting system causes them to try to untwist. Hairs from one strand are caught by the
other; the individual yarns twist about their own axes and consolidate the grasp of the
hairs from the other yarn. Meanwhile, the pair of strands twist about their common
axis to relieve the torque in the individual strands (Fig. 10.11(a)). There is a discernible
zero twist zone at each changeover and this zone increases as the yarns wrap around
each other in the separate twisted zones. The transfer of torque from the component
strands is reduced as the local ply twist increases and the system comes to equilibrium.
There is a resulting series of slightly extended zero twist zones, interleaved with ply
twisted sections of yarn (Fig. 10.11(b)).
10.7.2 Self-twist yarns
Let the length between changeovers be L (variable) and let the staple length of the
fiber be fixed at S. The zero twist zone is a weak link in the chain because any fiber
Handbook of yarn production
Strand 1 input
Strand 2 input
Rolls oscillate axially in
opposition to one another
ST yarn output
Fig. 10.10
ST yarn process
τ = τ1 + τ2
τ1
(a)
L
(b)
Relative yarn tenacity
272
1.0
S = Fiber length
0
L/ S
1.0
(c)
Relative yarn tenacity is a measure of yarn strength
relative to that which might be expected from using
a similar fiber in a ring spun yarn.
Fig. 10.11
ST yarns
Staple systems and modified yarn structures 273
that has an end in the zero twist zone contributes no strength. Assume that the fiber
ends are randomly distributed and let m be the number of fibers in the whole crosssection. It is necessary for L<S to provide fibers to transmit load across the weak
section. The number of ungripped fiber ends per cross-section in the zero twist zone
relative to those in the twisted sections will then be L/S and the number of loadbearing fibers is:
[1 – L/S]m
[10.1]
Ignoring fiber obliquity effects, the relative strength of the zero twist zone is:
[1 – (L/S)] × 100%
[10.2]
Growth of the zero twist zone has to be limited. Unless the strands are prevented from
rotating about their own axes, the torque in the strands tries to balance by removing
more twist from the twist changeover zones. This was briefly alluded to earlier.
Fortunately, long fibers catch on the other strand easily and the process of binding
fibers from the co-operating strand joins them and forms a torque stop. Lengths of
the zero twist zones are determined by this licking-in process. To get the necessary
fiber wraps, twist flows between zones. As the strands untwist, the length of the zero
twist zone increases. The strength of the zero twist zone varies with staple length in
the manner shown in Fig. 10.11(c). It will be seen that it is necessary to use a staple
length of several inches (unless the weak points are strengthened by some means) to
get a reasonable strength. Of the various means to stabilize the weak zones, a few will
be mentioned. Stomph [17] used intermittent air vortices created by switched nozzles
to false twist the individual strands before assembly at what would have been the
weak points. The system was complex and limited in speed to some 100 m/min; also
piecing was not simple. Morgan [18] sized the yarns with a water-soluble adhesive
before further processing, but this system did not achieve wide usage either. Air-jet
texturing can create lateral fiber migration and this is useful in locking the structure.
In practice, the ST process is confined to long-staple fibers.
10.7.3 The self-twist process
The self-twist (ST) process was alluded to in Section 3.4.2 and the discussion is
continued here. There is no conventional spindle or rotor and very much higher
processing speeds than normal can be used for ST spinning. Twisting and winding are
separated, with the result that large packages of unbroken yarn can be made. The
shuffling/twisting rollers are capable of inserting twist at extremely high rates (>10
× ring frame productivity) and the system is capable of producing yarn cheeses
containing up to 9 lb of yarn. To keep L small, high levels of alternating twists are
required. The yarns tend to be weak but they can be produced at very high speeds at
relatively low cost.
Phasing the component strands so that the zero twist portions of each strand no
longer coincide with the zero twist zone of the ply (Fig. 10.12) can alleviate the
problem in patterning. In practice, this is easily accomplished by making one of the
component strands take a longer path than the other, on its way from the twisting
rollers to the guide, J (Fig. 10.13). When the strands unite, one strand is out of phase
with the other. If the phasing is properly set, the patterning, as well as losses in
strength, reduced but it is still necessary to use a long-staple system. All commercial
ST yarns are phased.
274
Handbook of yarn production
Strand = zero twist
Ply = Zero twist
Fig. 10.12
Phased STT yarn
Strand 2 input
Strand 1 input
Guides
J
Rolls oscillate axially in
opposition to one another
Phased ST yarn output
Fig. 10.13
Phased STT yarn process
The self-twist process is rather simple. A normal machine contains a number of
production channels. Each channel comprises a roving supply package, a roller drafting
section with an oscillating front roll (i.e. shuffling/twisting rollers), a strand combination
system, and a take-up. The oscillating front roll provides the alternating twist as
already described. The machines have an exceptional productivity, but even the phased
product still causes some patterning in fabrics that makes them look streaky. The
simple process has a restricted market and much of the yarn is twisted to make STT
yarns as described in the next section.
10.8
Twisted self-twist yarns and processes
10.8.1 Twisted self-twist yarn
As already discussed, alternation of twist in ST yarns creates a streaky effect in a
fabric but it is possible to superimpose real twist sufficient to make the ply twist
unidirectional to minimize this effect. Such twisted self-twist yarns are known as
‘STT’ yarns. Nevertheless, even if real twist is inserted by two-for-one twisters, the
cost of twisting is relatively large compared to the cost of spinning and it is therefore
less attractive economically than it first appeared. Shaw [19] estimated in the early
1970s that the costs relative to the ring frame varied between 80 and 88% for counts
between 30s and 9s worsted respectively.
When STT yarns are made, the real twist added has to be high to minimize the
patterning. The result is that, with wool, the system produces a high twist, longstaple, plied yarn that has much of the character of a worsted yarn. A worsted yarn
is a relatively high twist product and therefore the STT system is still quite attractive
in this market.
Staple systems and modified yarn structures 275
10.8.2 Processes for twisted self-twist yarns
Two processes in series are needed to make STT yarns. The first is the manufacture
of the ST yarn and the second is the plying process. This involves transfer of yarn
packages between the processes, which increases the cost. It does, however, give
increased manufacturing flexibility which may be useful in a specialty market.
10.8.3 Processes using modifications of the ST process
One development is to self-twist a staple strand with a filament and then self-twist
this composite with another filament. This produces a staple core with filaments
wrapped in opposite directions on the outside. Fine filaments are not very visible and
the streaky effects can be minimized. The filaments add strength to the structure but
they also tend to increase the cost/lb. The trade name for this type of composite yarn
is ‘Selfil’.
A second development reported by Miao et al [20, 21] concerns STT yarns modified
by air-jet texturing. The air-jets interlace the yarns with the result that yarns tested
somewhere between 20% and 60% stronger than with non-interlaced yarns. Patterning
in the fabric was diminished as compared to normal ST yarns and fabrics.
References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
Götzfried, K. Process and Apparatus for Spinning Yarn, BP 825,776, 1959.
Pacholski, J. The Pneumatic Spinning Machine PF1 – Characteristics and Application, Theory
and Practice of the New Spinning Techniques, Instytut Wlokiennictwa, Poland, c 1980.
Stalder, H. The Possibilities and Limitations of Various Short-staple Spinning Systems,
UMIST Symposium on Tomorrow’s Yarns, Manchester, UK, 1984.
The Random House Dictionary, Random House, New York, 1966.
Morihashi, T and Murata, K K K. Japan, US Patent 4,183,202, 1980.
Nakahara, T and Murata, K K K. Japan, US Patent 4,142,354, 1979.
Hasegawa, J, Kawabata, S and Niimi, H. Toyoda, Japan, USP 4 434 611, 1984.
Yamana, M and Kubta, N, Murata, K K K. Japan, USP 4 107 911, 1978.
Jones, T. Application Note No 2, Oxford Lasers Ltd, Oxford, UK, 1990.
Oxenham, W and Basu, A. Effect of Jet Design on the Properties of Air-jet Spun Yarns, Text
Res J, 63, 11, 674–8, 1993.
Looney, F S. Optimierung des Luftspinnens durch den Einsatz von Dacron-Polyesterfasern,
Dornbirn, 1984.
Looney, F S. Engineering of Polyester Fibers for Modern Spinning Systems, Tomorrow’s
Yarns, (Ed Hearle J W S), UMIST Symposium, Manchester, UK, 1984.
Plate, D E A and Lappage, J. An Alternative Approach to Two-fold Weaving Yarn, Part 1, J
Text Inst, 73, 3, p 99, 1982.
Lorenz, R R C. Yarn Twisting, Text Prog, 16, 1/2, 1987.
Lamb, P R and Junghani, L. Drafting and Evenness of Wool Yarns Produced on the Plyfil,
Sirospun and Two-fold Systems, J Text Inst, 82, 4, p 514, 1991.
Xie, Y, Oxenham, W and Grosberg, P. A Study of the Strength of Wrapped Yarns, J Text. Inst,
77, 5, 1986.
Stomph, J G. Spinning Short-staple Fibers on the CSIRO-system, 4th Shirley Int Seminar,
The Hague, 1971.
Morgan, W V. The CSIRO Self-twist Yarn Spinning System, 4th Shirley Int. Seminar, The
Hague, Netherlands, Netherlands, 1971.
Shaw, J. New Methods of Yarn Production, Private communication, c 1974.
Miao, M and Lui, K K. Air-interlaced Self-twist Yarns, Text Res J, 188, 67, 3, 1997.
Miao, M and Soong, M C C. Air Interlaced Yarn Structure and Properties, Text Res J, 65,
433–40, 1995.
11
Quality and quality control
11.1
Quality
11.1.1 Definition of quality
Modern use of hyperbole has widened the meaning of quality to such an extent that
it is desirable to narrow it for the present purpose. The whole textile enterprise is
founded on bargaining between supplier and customer. Two of the most important
factors in the contract (explicit or implicit) are quality and price. Other factors, such
as delivery schedules, service, reputation, etc., also apply, but can be set aside for the
present argument. In textile technology, quality is often defined in terms of various
attributes of the fiber, yarn, or fabric, but this alone is insufficient. What forms the
basis of an acceptable bargain for a given product for one particular end use may not
be acceptable for another set of conditions. Consequently, quality may be defined as
a set of attributes for a product that fulfills the needs of a customer or user.
Interlinked with the physical characteristics of the product is the question of price.
Securing superior physical characteristics of the product often involves higher costs
for the supplier and this usually results in higher prices. Profit margins for the yarn
producer have to be sought in the differences between cost and price. The consumer
looks for the highest quality at the lowest price.
There are several aspects of quality control, some aimed at preventing difficulties
and some aimed at curing the cause of them. A routine of sampling, testing, and
adjusting is the standard method of day-to-day control. Also customer complaints or
difficulty reports often require test work. Consequently a quality and control department
is usually equipped with fiber, yarn, and fabric testing equipment.
11.1.2 Quality factors
Acceptable quality is determined by the user, and fabric makers are major users of
yarn. The desirable attributes for fabric may be classified as those related to fabric
appearance, fabric durability, and freedom from faults. Often, woven fabric durability
and yarn strength are found to be linked. Similarly, the ease with which the yarn can
Quality and quality control
277
be manipulated in making fabric is often related to the yarn strength, fault rate, and
hairiness. However, these relationships are not universal. For example, whilst warp
yarns for weaving are required to be strong, knitted fabrics have no great need for a
strong yarn. However, they both have a need for non-twist lively yarn with good
evenness of linear density, hairiness, and dye affinity.
The appearance category may be subdivided into evenness of linear density, hairiness,
coloration, light reflectivity, and refraction. (A minor segment of the market specializing
in novelty yarns may have different standards from those more generally practiced,
but space precludes further discussion of these.) Evenness is often expressed in terms
of variance, standard deviation, or CV of the attribute concerned; consequently there
is considerable discussion of these factors in this chapter. Fabric durability is a matter
of yarn strength, yarn structure, and fabric structure. Yarn strength is mostly a matter
of yarn structure and fiber strength. These matters become more complex when blend
yarns are used, especially those in which the fiber properties of the constituents differ
greatly. Freedom from yarn faults involves not only minimizing the production of
faults during spinning but also the removal of them in winding. Furthermore, removal
of a fault requires the joining of cut ends and the join itself is sometimes an unacceptable
fault. Yarn processability is important, not only to the fabric maker but also to the
yarn maker himself. Obviously, a weak yarn is more difficult to process than a strong
one. Other factors also apply, and these include twist liveliness, yarn hairiness, residual
yarn fault level, and yarn package construction. Deterioration in any of these factors
can cause difficulties in yarn manufacture and in quality of the product.
11.1.3 Analysis of customer complaints
Since the standard of quality is set in the marketplace, and since the standard changes
from time to time, it behoves the spinner to keep abreast of what technical properties
the market is demanding and how the product mixes of his or her company meet that
demand.
Again, there is no absolute definition of what the demand is, and considerable skill
is needed in interpretation of the data. The customer complaint level gives a good
window on the technical requirements but it varies according to the state of the
economy. When the economy is booming, complaints lessen and there is a danger of
complacency among those responsible for the quality control system. When the
economy declines, complaints mount, and unjustified complaints are mixed with the
justifiable ones. Nevertheless, analysis of the customer complaints is a prime tool for
keeping track of the quality levels. Technical analyses are also important because, not
only do they permit the solving of problems, but they can provide an information
channel to the marketing people. Thus, the testing facility is an important part of the
business.
A typical spectrum of complaints is shown in Fig. 11.1, but the spectra vary
according to the nature of the businesses. Nevertheless, the sample can be used to
make several points. Mundane affairs, such as shipping, can sometimes be even more
important than a technical issue such as the CV of linear density. The sample quoted
in Fig. 11.1 was for fine staple yarns and it might be noted that yarn appearance and
winding ranked at the top of the list. Neither of these categories uses a single
measurement as a criterion. Rather, the judgment is made using a complex list of
factors. As a contrast, filament yarns have different criteria and complaints range
more in the field of polymer morphology than in evenness of the product. Changes
278
Handbook of yarn production
Customer complaints/year
20
Sample spectrum only
15
10
5
Fig. 11.1
Yarn weak points
Yarn count
Yarn strength
Yarn hairiness
CV of mass
Foreign fiber
Package build
Shipping
Winding
Yarn appearance
0
Spectrum of complaints
in mechanical and thermal stress history are very important in this latter field since
stress history determines the dyeing performance of these types of yarns.
Winding complaints are often derived from unsatisfactory conditions in the earlier
processing. Thus, a high volume of complaints labeled under ‘winding’ might stem
from bad spinning, which might, in turn, arise from poor spinning preparation. Yarn
appearance usually includes yarn faults such as neps, thick and thin spots, etc. It also
includes some factors that are difficult to quantify by laboratory measurements. The
importance is illustrated by the fact that more than 16.5% of all fabric faults in
shirting, in one set of market data, were from spinning faults.
The item ‘foreign fiber’ usually comes from improper stripping of the bale coverings
in the process of laying down the bales in preparation for the next offtake run. It is
not easy to handle a bale of fiber after the straps have been removed. Consequently,
the operators have to develop techniques of removing every last strand of bale wrapping
from the underside of the bales. Regardless of whether the bale wrapping is jute,
polypropylene or some other material different from the fiber within the bale, the
foreign fibers show up after finishing the fabrics. Sometimes it is due to differences
in dye affinity, sometimes to differences in fiber size and color.
11.2
Quality control
11.2.1 Fiber quality control
Testing is a very important part of quality control and needs more space than can be
allocated in this chapter. Most of the data presented in this section were gathered in
the 1990s but they are subject to change as developments of equipment and techniques
improve. Many of the various quality factors have improved by roughly 0.6% per year
in the last half a century and the trend is likely to continue in the twenty-first century.
The Uster Corporation periodically issues a very comprehensive statistical analysis
of worldwide data under the title Uster Statistics [1] and the reader is referred to the
current issue at the time of need for information.
A well-founded quality control program should carry out tests in a controlled
Quality and quality control
279
atmosphere and all samples should be conditioned by immersion in that atmosphere
for an adequate time. Further discussion is given in Appendix 4.
Routine tests require well-founded sampling plans. The raw materials should be
sampled regularly but the schedule depends on the fiber and the variability thereof.
Generally speaking, there is less testing of man-made fibers in the mill as compared
to that used for natural fibers. In this case, greater reliance is placed by the mill on
the fiber maker for help in quality matters than is the case with natural fibers.
Consequently, remarks in this regard will be mostly confined to natural fibers.
Labeling and sampling varies according to the type of fiber. Modern practice with
cotton is that every bale is labeled with the normal fiber parameters, but that does not
necessarily mean that every bale has been tested. There is a growing practice of
module averaging in which a fairly large number of samples is measured from a
module and the averages are assigned to every bale taken from that module. Debate
continues whether it is better than testing every bale with only two or three samples.
Some yarn makers use HVI or other mass testing lines to measure every bale,1 to
assist them in preparing optimum cotton blends. These testing lines carry out a
battery of standard tests on fiber (usually cotton) in a continuous fashion. HVI stands
for ‘high volume instrument’ used in testing cotton, and it is a proprietary name. With
wool, the fibers may be in bulk, and sampling might require core-boring tools. The
mass is sampled randomly; the samples are subdivided, doubled with other samples
from the same mass, and then subdivided again before testing. Sometimes there are
more than two subdivisions and doubling stages in the sampling process. The
International Wool Testing Organization (IWTO) specifies that there should be at
least 100 test zones. With bast fibers, at least 20 bunches are selected at random, and
a strick is removed from each and tested. The strick is divided lengthwise, one
portion is discarded, and the remainder is separated into tip and root portions. These
are halved repeatedly as necessary, the tip and root samples being kept separate.
Composite samples of both tip and root are tested. Even man-made fibers are tested
by some with the dividing and doubling techniques.
Of the fiber attributes, length is often considered the most important. Specific
ways of testing are given in Appendix 4 but it is thought worthwhile to give an
example of how processing and testing can influence the results. Normally, fist-sized
samples of fibers from the bales are brought to the laboratory and are conditioned
before testing. A clamp is used to secure and withdraw a sub-sample of fiber from
each main sample submitted. The fibers protruding from the clamp are called a
‘beard’ and the fibers actually tested are in this beard. To measure length, it is
necessary to straighten the fibers by some sort of combing action. The number of
fibers in a cross-section of the beard is determined by light penetration through, or
electrical capacity of, a small ‘slice’ of the beard running parallel to the clamp. The
measuring head is traversed perpendicular to the clamp to a position where it measures
50% of the signal it had recorded at the clamp. The position of measuring head is then
taken as a measure of fiber length. There are different ways of testing, which cannot
be further discussed here.
The combing removes loose fibers and changes the fiber distribution. Compare a
1 Choosing only one or two samples per bale provides little or no guidance about the within-bale
variance of the fiber attributes measured. These variances are often of the same order of magnitude
as the between-bale values measured. Consequently, one of the dimensions needed for total
control is missing.
280
Handbook of yarn production
Test zone
Fibers
(b)
(a) Clamp
y
Fibers
Clamp
x
x
Histograms
(d)
(c)
Fig. 11.2
Fiber length sampling
theoretical result before combing with one after. Figure 11.2(a) shows a population of
fibers, some of which are clamped and some are just free of the clamp. The latter are
shown in gray whereas the clamped fibers are shown in black. A real sample would
contain hundreds, if not thousands, of fibers but for clarity the diagram shows only
a few. If all the fibers in the zone are counted, the histogram of length in the sample
is as shown in diagram (b). If the unclamped fibers that were shown in gray are
removed as in (c), the histogram changes as shown at (d). Even if the statistical
frequencies are normalized to 100%, the histograms still differ. The latter is a lengthbiased sample, which differs from reality. A third form of sampling is when the fibers
along the line xx are counted as a function of y. This is called a tuft curve. The results
can be expressed as histograms or cumulative frequency curves. Thus, the method of
testing and the history of the material can strongly affect the result. This is important
when the results are the basis of decisions, especially if the methods used by the
supplier and the supplied are not co-ordinated.
In HVI testing, most of the loose fiber is removed from a beard and a 50% span
length is measured. This is the value of y when the number of fibers along xx is 50%
of the total in the clamp. The 50% span length is about 0.58 times the classers’
length.2 Detailed distributions are given in the Uster Statistics 1997 [1]. Mean shortfiber content is shown to be unaffected by processing but the amount of trash steadily
decreases with processing. This is not to say that the CV of short-fiber content is
unchanged; as will be shown later, there can be significant differences. Again, according
to Uster, the short-fiber content by weight can vary from 14% down to 6% for the
best 5% of production.
A second important fiber attribute is fiber fineness. With wool, fiber fineness is
measured by fiber diameter expressed in microns, whereas with cotton, the fineness
is usually expressed by the micronaire index. Micronaire is a measure of the permeability
of a fiber wad of a defined size when it is mounted in a defined chamber. While it is
true that micronaire is related to fiber fineness, it is also related to the maturity of the
cotton; nevertheless it is an industry standard. For many sorts of American cotton,
there is little variation in the micronaire/fiber fineness ratio with fiber length. However,
there are some high micronaire cottons that display up to 30% difference between
short and long fibers. If long-staple cottons are excluded, the differences can usually
2 The classers’ length is one obtained by a manual method now largely obsolete.
Quality and quality control
281
be ignored for fibers between 3 and 4.5 micronaire; in other words the micronaire
value is a reasonable measure of fiber fineness within the quoted range.
A third fiber factor with cotton is yellowness, as measured by the coefficient ‘+b’
[1, 2]. The value varies from 7.9 to about 11 for short cottons and from 8 to about 13
for long cottons, the higher figures being more yellow than the low ones. The yellowness
comes from natural dyes in the fiber and from yellow-tinged wax coatings. It is
desirable that the yellowness should be low to reduce the scouring and bleaching of
the fabric that might be necessary; above all, it is desirable that the figure should be
uniform throughout the product. Preliminary measurements [3] suggest that there
might be as much as 5% difference between adjacent lengths of sliver. Generally, few
measurements have been made of this variability, but perhaps more attention should
be paid to it.
A fourth factor with cotton is the amount of trash present. A typical HVI measurement
relates to the surface area of a bale sample occupied by dark colored trash. According
to the Uster Statistics [1], the trash surface area reduces as the fiber length increases.
For short cottons, the worst trash counts can change from a value of 2% to 1.5% as
the fiber length changes from about 1.0 to 1.2 inches. The best case has a count of
under 0.1%, irrespective of fiber length within the quoted range. The long cottons
show a similar pattern, although the values are lower.
Other important factors are nep and short-fiber content. With short cottons, the
patterns of both neps and short-fiber content with respect to fiber length are similar.
Long cottons show fewer neps and lower short-fiber contents than short cottons.
11.2.2 Quality control of intermediate products
Intermediate products such as sliver, roving, etc. are tested on a routine basis; the
items tested depend on the business. Commonly, linear density of the intermediate
material, such as sliver or roving, is measured daily at each step; yarn strength, yarn
hairiness, neps in the card web, and yarn fault levels are also checked daily. However,
even strict periodic sampling may give erroneous results if there is a periodic variation
from a prior process that is slightly different from the sampling frequency. The two
frequencies are said to beat against each other and they produce a beat frequency that
shows up in the result. Where exploratory tests are being used to diagnose a problem,
a viable experimental plan is required and that involves a knowledge of the variables
likely to cause the problem. When long-staple sliver samples have to be transported
between plants or units, it is desirable that sliver be twisted to prevent disturbance of
the structure and fiber distribution (for wool tops between 15 and 30 ktex, Anderson
[4] recommends 20 turns/m).
In assessing the results of testing, it has to be realized that the total variance3 of
random errors measured is the sum of the variance between the bales, zones, or other
large divisions and the variance within them. The between-zone variance is possibly
due to fiber acquisition policy whereas the within-zone category is a micro-variation
in the supply, often caused by processing. This distinction is sometimes helpful in
seeking to reduce the variance in properties.4
3 Variance is the square of the standard deviation.
4 Standard deviations have to be weighted to take into account the number of samples taken and
(Standard Error)2 = s2/ms + s 2 /mz.
282
Handbook of yarn production
11.2.3 General yarn defects
Defects are usually regarded as single, random deviations from the normal parameters
describing the yarn. For example, a single fairly long thick place (or ‘slub’) would be
called a defect, whereas a systematic series of thick and thin places would usually be
classified as irregularity or unevenness. Unevenness will be described later. Sometimes
slubs appear periodically; they are recognizable by their torpedo shape.
Defects can occur in either staple or filament yarns. In filament yarns, many
defects arise from differences in morphology of the polymer rather than from differences
in linear density. (Morphology is a term used to describe the molecular structure of
polymers.) Changes in morphology result in alterations in dye affinity that have a
powerful effect in the category of yarn and fabric appearance. However, married
fibers, drips, and debris also cause problems. The dominant defects in staple yarns
arise from processing the fibers and interactions between procesing and the fibers.
The possibilities of producing unacceptable yarn for staple yarns are greater than for
filament yarns because a large percentage of staple yarns contain natural fibers, with
their inherently variable sets of characteristics. Also natural fibers are associated with
non-fibrous materials, which have to be removed. This is not to say that filament
yarns are without problems. Complaints arising from filament defects are a matter
between the spinner and his or her fiber supplier whereas natural fibers are subjected
to some uncontrollable influences such as weather. For these reasons, discussion will
be centered on staple yarns.
As mentioned, yarn defects can be caused by a variety of circumstances. Some
machine errors tend to be organized, but others occur at random. The organized
effects, such as those due to eccentricities and drafting waves, have already been
discussed. Random errors, such as the production of slubs, piecings, corkscrews,
crackers, etc., have not yet been discussed (see Fig. 11.3). Many of the fault types (a)
through (d) are caused by drafting. Drafting under too dry conditions can cause static
electricity to be generated by the sliding fibers, with the result shown in diagram (c).
Balls of fiber on the yarn caused by accumulations of lint on the traveler (d) are often
composed of short-fiber debris from drafting. Occasional long fibers bridge the nip
lines in the drafting system and disrupt the process (e). Loose fly, captured by the
Name (cause)
(a)
Thin spot (drafting)
(b)
Slub (drafting)
(c)
Fishes (static)
(d)
Ball (traveler)
(e)
Cracker (long fiber)
(f)
Spun-in fly
(g)
Piecing
(h)
Nep (ginning & carding)
Fig. 11.3
Some fault types
Quality and quality control
283
Cumulative defects/unit
length arbitrary units
Total defects/unit length
arbitrary units
yarn between the front drafting rolls and the pigtail guide (f), is usually ‘spun in’ at
only one end [5], whereas fly deposited before that is held more firmly. A piecing is
the result of an end-break (g) and neps (h) have already been discussed. If the best
mills now are compared to those described by Thomason in 1971 [6], one can see
reasons for some of the improvements by merely walking around and noting the
vastly improved state of cleanliness. Some results typical of the ring spinning industry
are shown in Fig. 11.4(a) [2]. Following the discussion in Section 3.7.7, where the
known increase in difficulty in drafting at high ratios is discussed, it should be no
surprise to find that the fault rate is a function of the draft ratio. Uster Tester data [1]
relevant to thin spots in the yarn show that an average carded cotton yarn spinner
might experience ranges from 1/km to 150/km dependent on counts between 16s (38
tex) to 45s (13 tex) respectively. The corresponding thick spots range from 120/km to
800/km. Not only are the thick spots more prevalent but they can cause trouble in
later processes unless removed. Defects have a range of thicknesses and lengths that
are classified accordingly. The Uster Classimat [7], an apparatus for classifying yarn
defects by length and thickness, has four length classifications, labeled A to D, which
range from 1 mm to 40 mm. It also has 4 thickness classes, labeled 1 to 4, each
centered at various levels between 100 and 400% of normal yarn linear density. Often
there are relationships between these thicknesses and lengths (e.g. Fig. 11.4(b)).
There are fewer large defects than smaller ones and there are fewer long defects than
shorter ones, as might be expected. A profile of combinations of length and thickness
is set to give limiting settings for defect removal in winding. This usually includes a
protocol at various levels in the A3, B2, C1, and other categories. Multiple winding
of yarn to remove defects can also damage it. Polyester/wool and acrylic fibers are
especially vulnerable to such damage.
In staple spinning, irregular faults can be produced by improper maintenance,
machine settings, and raw material supply. For example, too close a roll setting in a
drafting system can lead to fiber breakage, slub creation, and derivative faults such
as traveler accumulations of fly. Wear of machine parts occurs after a period of use
and this wear frequently leads to the production of defective yarn. An unclean or
improperly conditioned atmosphere can cause problems. Failure to remove trash in
opening and carding can lead to the production of yarn faults, especially for trash
over a certain small size. One can often discriminate between failure of the cleaning
machines in the opening line to remove trash and the failure of the card to remove it,
by inspecting for trash particles embedded in the card flats. Poor cleaning in the
0
(a)
20
Draft
40
Short
Long
0
100
200
300
Relative size of thick spot
(b)
[(n′ – n) × 100%]/n
n ′ = linear density of thick part
n = linear density of normal part
Fig. 11.4 Defects in yarns
284
Handbook of yarn production
opening line results in significant deposits of trash in the wire of the flats. The
production of neps is frequently a significant quality control problem; controls at
carding help in this respect.
11.2.4 Staple yarn defects arising from the fiber
Defects inherent in the supply of natural fibers arise from the presence of non-fibrous
material and defective fibers. These may be categorized as:
1
2
3
For cotton: immature fibers, neps, trash, etc.
For wool: grease, suint, vegetable matter, etc.
For man-made fibers: concentrations of finish, oligomers, undrawn or improperly
drawn segments of fiber, etc.
Other fibers have foreign or unwanted matter in their supply too.
These materials adversely affect processing, which, in turn, produces error. But
here we consider the direct effects on such properties as dye uptake, fiber reflectance,
and appearance of the final product.
Differences in fiber color can also produce disturbing effects leading to complaints.
The differences can be from batch to batch, there can be differences within a yarn lot,
or there can be differences within a package of yarn. These differences are categories
of yarn length (which range from a few yards to thousands, or even millions of yards)
and the fabric faults they produce fall into the categories of barré and streaks. Figure
11.5 shows two examples of color variation in fiber yellowness. The top diagram
shows the variation in card sliver that was made from a single bale of cotton. The
inch-to-inch measurements5 show a variation range of about 5%, despite being taken
from source that is often treated (wrongly) as invariable. The bottom diagram shows
the variation in a sample of commercial combed sliver in which the number of sliver
doublings involved was greater than 3000. Despite the doublings, a significant error
is visible and the wavelength of the whole error cycle was probably in the hundreds
of yards. Such a variation could have led to barré problems in greige fabric.
Nep varies from processing stage to processing stage. According to Uster [1],
the nep level in the bales (≈ 100 to 900 nep/g) is slightly lower than the level in
the fiber approaching the card. This level is then reduced in carding to the range
25–300 nep/g. For the longer cottons used for combed yarns the nep counts in the
bales are less and they can be reduced to the range 7–80 nep/g after combing. The
neps in this form of measurement are of an absolute minimum size. The actual
reductions are determined by the fiber as well as the design, settings, and maintenance
of the machines. An average ring spinner produced carded yarn between 6s (100 tex)
and 45s (13 tex) with nep counts of 150 and 50 nep/g respectively (AFIS Data [1]).
In an Uster tester type of measurement, a nep is defined as a very short fault of more
than 200% of the yarn diameter. Consequently, with this type of measurement, the
size of the neps recorded, vary with the yarn count. Data from the Uster tester [1]
suggest that average ring spinners produced nep levels in the range from 25 nep/km
5 It is not possible to measure such samples on an HVI instrument and therefore these measurements
were made on a flat-bed scanner. Adobe ‘Photoshop’ software was used as a photometer applied
to color images of sliver, the images were converted to the CMYK mode, and the yellow
separation was used. The percentage of yellow represents the color depth and is related to the +b
value normally used in textiles.
Quality and quality control
285
% Yellow
15
10
5
Card sliver from a single bale
0
0
5
10
% Yellow
15
10
5
Commercial combed sliver
0
0
20
40
60
Length along the sliver (yards)
Fig. 11.5
80
Variation in fiber color
(≈ 0.3 nep/g) at about 6s cotton count to 1300 nep/km (≈ 100 nep/g) at 45s cotton
count. A 45s combed yarn gives an average nep count of about 12 nep/g (it would be
rare indeed to make a 6s combed yarn). Thus it can be seen that care is needed in
interpreting results. In rotor spinning, the corresponding AFIS figure is about 200
nep/g irrespective of count. Thus, rotor spinning produces a slightly inferior yarn as
far as nep is concerned. This may be due to nep created in the combing roll of the
rotor spinning machine. In cotton spinning, neps are often associated with other
defects, for example as in Fig. 11.6(a). To emphasize the importance of the quality
control regime in a mill, consider Fig. 11.6(b), which shows the results of tests in two
separate mills spinning different counts of cotton yarn. A 50s yarn usually shows a
different nep count from a 30s, but this is not the issue in this case. Rather, one mill
Mill A: y = – 4.14x + 113
Mill B: y = 0.42x + 28
2
y = –0.006x + 1.446x + 10.33
150
r 2 = 0.898
r = correlation coefficient
100
Nep/1000 yd
Nep/1000 yd
150
50
100
50s
50
30s
0
0
0
25
50
75
Thick spots/km
(a)
100
0
Fig. 11.6
Nep control
10
Months
(b)
20
286
Handbook of yarn production
had worked to reduce neps and the graph shows the result. The other mill had a
reasonably good nep count for the end use and, unfortunately, saw little need to work
hard at the problem. The example is intended to show the value of progressively
plotting fault production.
Some data relating to yarn faults in various types of yarns are shown in Fig. 11.7.
The yarns are of average quality. All the faults shown decline in number as the linear
density increases (i.e. the count decreases). Worsted yarns show up poorly in thin
spots whereas 100% carded cotton yarns show up badly in thick spots. It is, perhaps,
not surprising to find the number of thin spots for worsted yarns to be high when it
is remembered that wool is more variable in length and fineness than cotton. Also, it
is not surprising to find that combed cotton yarns perform well in this respect. It is
interesting to see that the nep performance for wool is relatively good, whereas the
performance of carded cotton is relatively poor. Presumably this is because cotton is
so much finer than wool. It is a little disappointing to see the relatively poor nep
performance of air-jet yarns; perhaps this is due to the extra drafting.
Trash, dust, and visible foreign matter are progressively reduced by processing
and the levels in combed yarn can be reduced to the order of 1% of the values
pertaining to the bale material. The best and worst spinners of carded ring spun yarn
produce 0.1 and over 10 trash particles/g, respectively, at 50s count; the figures for
6s yarn are approximately 8 and 30, respectively. Thus the quality of the processing
is seen to play a very significant part in the quality of the product.
The use of waste fiber affects quality. Before regulations restrained yarn makers
by making truthful labeling mandatory, more mixed waste was used than now. Thus,
we had shoddy in the wool trade, recycling of blend fibers of indeterminate blend
ratios, filaments and fibers made from recycled polymers, and so on, but such practices
are rarer today. The name ‘shoddy’ is now synonymous with poor quality. Comber
noils are sometimes recycled within a mill but it is also a frequent practice to sell the
noil to yarn makers who make lower grade products. For these latter people, the noil
Thick spots/km
K1
100
K2
A
10
A
1000
W
K3
K1
K2
100
10
C1 & C2
W
C2 C1
1
1
20
20
50
100
200
Yarn linear density, log scale (tex)
K3
100
C1
A
C2
K2
10
W
1
50
100
200
W = Worsted
K1 = Carded, 50/50 P/C, rotor
K2 = Carded, cotton, rotor
A = Combed, 50/50 P/C, air-jet
K3 = Carded cotton, ring
C1 = Combed, 65/35 P/C, ring
C2 = Combed cotton, ring
K1
20
50
100
200
Yarn linear density, log scale (tex)
Fig. 11.7
K3
Yarn linear density, log scale (tex)
1000
Nep/km
Thin spots/km
1000
Fault rates for average yarns
Quality and quality control
287
is a relatively cheap raw material. Undercard and flat waste are sometimes recycled,
but much of it is disposed of. Pneumafil waste (see Section 5.11.3) is nearly always
recycled within the staple mill. Intense competition and pressure on prices make
prudent recycling a necessity. As in Fig. 11.8, the amount of pneumafil generated
with cotton yarns is substantial and it rises with count. Recycling has to be carried
out with care.
In filament/staple processing there is less possibility of recycling the waste internally.
11.2.5 Staple yarn defects arising from processing
The analysis in this section will be organized from winding back through the processes
until we reach the bale laydown. Many yarn faults are removed during the clearing
operation and are replaced by standard knots or splices. Not only does this make it
possible for faults to be produced and then be removed without the manager being
aware of the fact, but it calls into question whether the yarn should be tested before
or after clearing. Normal practice is to test samples after piecing. Study of the
piecings from a given operator or machine will quickly show a characteristic appearance.
This sometimes enables the source of a problem to be traced.
Every time an end breaks, the process has to be restarted by piecing, and these
piecings are always imperfect, even if they are acceptable. If the purpose of testing is
to detect the source and frequency of fault production, then there might be an incentive
to test the yarn before clearing. It would then become necessary to sample the thousands
of ring bobbins or equivalent packages entering the winding process.
If the purpose of testing is to protect the customer, it is necessary to sample the
wound packages destined for the customer. These latter packages contain the contents
of the input bobbins minus the portions of yarn removed, but with the added knots or
splices. It is not possible to regard these final packages as single specimens; in reality
each one is a series of yarns from almost randomly chosen bobbins of yarn. However,
with linked winding systems, the order of the component yarns on the cheese or cone
is usually the same as that of the rovings on the ring frame, which can be turned to
advantage. The amount of testing required to truly sample the material delivered is
much higher than normally realized. This is especially so if there are large variances
between individual bobbins. Remember that, as the CV reduces, the required sampling
frequency is also reduced.
The residual fault rate describes the yarn after clearing. It depends not only on the
original processing during yarn manufacture but also on the setting of the clearing
and splicing mechanisms. Thus, test results may be ambiguous. The levels at which
the clearers are set determine the size of defect removed, and the setting of the splicer
Pneumafil (%)
6
y = 0.071x – 0.754
4
2
0
20
Fig. 11.8
40
60
Yarn count Ne
80
Pneumafil production rates
288
Handbook of yarn production
determines whether the splice is acceptable or not. The more defects removed, the
greater is the chance of an unacceptable yarn. A frequent problem in this latter regard
is the production of long fiber tails to the splice. Such tails interfere with the processes
that follow and these faults should be avoided.
Defects can be caused in roller drafting as well as the normal set of errors already
described. Wrong ratch settings, worn guides, eccentric spindles, etc., might cause
fiber breakages or other fiber disturbances. Nicked, fouled, or damaged pigtail guides,
balloon control rings, travelers, rings, flyer grommets, or the like can also lead to
defects in the product. The nicks become more prevalent after certain abrasive manmade fibers have been used on the particular equipment. Often these machine defects
cause episodes of hairiness not easy to detect in the yarn but which become very
troublesome in later processes.
Apart from locally irregular drafting, fiber debris may be discharged into the air.
Since the highest draft is at the spinning frame, this is a good place to look for the
sources of defects. Concentrations of fly from this or other sources may then become
wrapped around the yarn to produce faults. Slubs and fishes are usually created by
electrification or by faulty settings of the drafting system. It might be noticed that the
defect level is a function of draft, all other things being equal. A badly arranged or
maintained traveling cleaner can blow concentrations of fiber onto the yarn, roving
or other strand. Raw materials that contain an excess of short fiber might lead to an
undesirable discharge of fly into the atmosphere. If the atmosphere is such as to
encourage fiber electrification, this fly may concentrate into tiny clumps which deposit
on the yarn during manufacture. Alternatively, the fly can be deposited on material
being stored in the workplace. Undrawn or married fibers, particles of foreign material,
irregular fibers and the like can temporarily interfere with the drafting process and
produce slubs. Deposits of fibers on the traveler or elsewhere may cause defects if the
deposits are suddenly licked into the product stream. Also they can cause the endsdown rate to increase. A sharp look-out for such accumulations is required.
When spinning blends of natural and man-made fibers, there can be accumulations
of fiber finish that become particularly troublesome on the balloon control rings,
main rings, and travelers. The deposits are often white and powdery, which explains
the common description of it as ‘snow’. They might vary in quantity from one fiber
maker to another or from one fiber merge to another, and some are very easy to
remove by washing, but economics do not allow for a washing operation of machine
parts.
There is an intermediate case where too light a roll pressure creates a string or a
series of defects rather than a classical case of irregularity. The effects of too light a
roll pressure can be magnified if the roving is fairly highly twisted and is irregular.
The hard ends described earlier can cause outbreaks of such strings of defects. All
these various examples have in common an irregularity of occurrence, but an experienced
eye can usually recognize the probable source.
11.2.6 Staple yarn defects arising from air conditioning
A cause of defects is the entrapment of accumulations of fly on the roving or yarn,
either at the ring or roving frame. Spun-in fly can usually be associated with poor
cleaning procedures or equipment. Every time an end breaks in a roving frame, the
broken end lashes the neighboring machine parts and creates a ‘snow storm’ until the
frame can stop. Inertia of the machine prevents an immediate cessation of motion.
Quality and quality control
289
Therefore, even if the machine is switched off immediately after the break, it continues
to broadcast fiber until it stops. There is a ‘run-down’ time. Apart from this, there is
a steady discharge of fibers from the roving and drawframes as they work normally.
This discharge often results in the creation of very loose fluffy rolls of fiber on the
floor of the mill; these often escape the patrolling cleaners, some parts of them
becoming airborne again and migrating to other areas. Also, concentrations of fiber
finish that accumulate on machine parts can cause irregular faults. If any of the
accumulations fall into crucial operating zones of the machines, faults are created.
In ring spinning, the traveler sometimes pushes trapped fly into ball-like shapes
and this type of defect can sometimes be associated with a high incidence of traveler
fouling. It is vitally important to keep the critical areas free from lint accumulations,
which is no easy matter. For example, a small fraction of a percent of the fiber
passing through a frame is liberated into the air to form fly. This may not sound much,
but in fact it represents hundreds of pounds of fly being deposited every day. If the
deposition rate is 0.02% of the fiber flow in a mill processing 2000 lb/hr, the deposition
rate is over 60 lb of fly/week. If you do not believe it, fix a wire mesh in front of (say)
one section of rings and you will find that, in an hour, several milligrams of dirty
fiber will be collected. At the ring rail level, the fibers will be on the side of the mesh
nearest the rings. At the level of the tops of the bobbins, the deposits will be on the
outside. The rotating bobbins act as a pump that sucks air from the room into the
balloon space. If the air is dirty (it usually is), fibers and contaminants are pulled over
the freshly made yarn being wound onto the bobbin. Adequate cleaning is vital and
this involves the use of traveling cleaners as described earlier. The main air conditioning
should also have adequate cleaning capabilities. Management must also consider the
sources of lint, because it is better to eliminate as many of the emissions as possible
rather than clean them up afterwards. On the supervisor’s tour, a sharp look-out for
accumulations of fly around light fittings, ceilings, and roof fittings is very helpful
in this respect. A study of the airstreams within a mill can help to determine if fly is
being transferred from a seemingly non-critical area.
11.2.7 Defects in fabric
The ultimate product is usually in fabric form, and a judgment of quality is normally
made on the number of defects per square yard. Frequently, demerit points are assigned
according to the length, diameter, and type of defect. For that reason, many spinners
knit samples of yarn to test for defects and barré. In weaving, there is a requirement
for adequate yarn strength and a lack of any defects that would cause end-breaks in
beaming or weaving. For example, consider a warp beam with 3000 ends of yarn
that has a defect rate of ‘only’ 10 per 1000 yards. On average, there will be 3000 ×
10/1000 = 30 faults per yard of warp. This is unacceptable. The faults can give
problems in at least three ways: (a) the end-breakage rate in beaming and weaving is
increased, (b) the fabric is degraded because of the yarn faults, and (c) the fabric is
further degraded because of the weaving faults due to the increased breakage mentioned
in (a). Figure 11.9 shows measurements of yarn strength made just after spinning and
it will be seen that the strength is not constant. As mentioned elsewhere, end-breaks
in spinning occur when the yarn tension exceeds the strength of a weak link in the
yarn in the process line. A similar philosophy applies to weaving but now the weak
spots concerned are in the wound yarn. Some faults have been removed but the yarn
might have been strained by high tensions or damaged, and a new crop of weak spots
Handbook of yarn production
Yarn strength
290
x
x
The line xx represents the applied load.
The portions of the yarn strength curve
shown in black represent potential yarn
breakages.
Time
Fig. 11.9 Variation in yarn strength
might now exist. Straining the yarn reduces the work needed for rupture and the
ability to weave well. Thus, the picture given by yarn testing may well be too optimistic
and the weaving performance might be worse than predicted. To demonstrate the
impact of yarn faults, let us consider a square weave fabric made from the same yarn
for both warp and filling. If the yarn diameter is proportional to (1/√Ne), where Ne is
the yarn count, then the end or pick density is (k√Ne)/CF per unit length. The factor
k includes not only the constant relating yarn diameter to (1/√Ne) but also takes into
account the weaving crimp: CF is the cover factor, which is defined as (area covered
by one or more yarns)/(area of the fabric). The length of yarn (L) in a square yard of
fabric is:
L ≈ [# ends/yd + # picks/yd)
L ≈ (2k/√Ne)(CF)
[11.1]
If the defect frequency is f defects/yard, the average number of defects/square yard is
Lf. If the fabric has a different construction (say, woven or knitted), we may replace
2k/CF by a different factor (K), which takes into account the difference in structure.
Number of defects/sq yd = Kf √Ne
[11.2]
In ring spinning, the fault rate increases with draft, as shown earlier in Fig. 11.4; the
major draft occurs in spinning and the largest number of defects arise there. The
number of defects can, in practice, be related to the yarn count and the quality of the
spinning operation. The cover factor does not vary greatly within a given class of
fabric; therefore, fabric defect rates for a given class of fabric are affected mostly by
changes in √Ne.
11.2.8 Testing for yarn defects
Usually, yarn defects are classified by length, local linear density, and fault frequency.
A typical distribution is shown in Fig. 11.4(b). It will be seen that short defects are
more common than long ones. Generally, a series of combinations of length and
linear density of the faults is used to determine whether the fault should be removed.
The choice of these combinations depends on the market involved and the experience
of the user. The amount of yarn that must be tested depends on the fault frequency.
With a fixed length of yarn (say 100 000 yd), there are two opposite dangers. On the
one hand there may be costly over-testing of a yarn, but on the other hand there may
be a risk of having tested an unrepresentative sample. In consequence of the need for
long samples, there is considerable incentive to use online monitoring.
Quality and quality control
291
11.2.9 Monitoring
An example of the interplay between economics and technical advancement is the
use of online monitoring at the ring frame. Because of the low output of a ring
spindle, the amount of capital expense that can be justified is limited. Monitoring
every spindle means that thousands of measurement points are involved and the
provision of a transducer of any great complexity would be too expensive. A solution
is to have a patrolling sensor that detects the presence or absence of the thread leaving
the drafting system. One such device is carried by a patrolling cleaner and another
uses a transport system mounted on the ring rail. The direct advantage is that those
end-breaks are signaled immediately they occur and this permits closer management
of the repairs. An indirect advantage is that this allows analysis of end-break patterns
and becomes a means to study regional effects of the air conditioning, pneumafil
settings, and machine maintenance. The system can have a beneficial impact on the
processing costs and product quality but the capital cost is still high and the advantages
have to be weighed carefully against the costs of installation.
Another interesting development is the roving stop mechanism. The idea is to
clamp the roving as it enters the drafting system when the end breaks. This reduces
the amount of pneumafil greatly. When the end is repaired, the whole drafting system
for the particular spindle has to be rethreaded. This can be automated. It has been
seen to be very effective from a quality point of view and the cost can be moderate.
Adaptations have been made to use the electrical signals that operate the clamps for
spindle monitoring.
Monitoring includes searching for defects. For example, it is common in rotor
spinning and filament production systems to use online recording of the passage of
defects in the flowing material. This is especially important because the product is
rarely rewound before use in manufacturing the fabric. Such strategies move the
testing from the laboratories to the production machines themselves.
11.3
Yarn evenness
11.3.1 General
Yarn evenness is made up of several components. Repeating what has been said
elsewhere, errors come from mechanical problems or from fiber flow problems.
Either may produce short- or long-term errors.
11.3.2 Harmonic irregularity due to mechanical faults
The class of mechanically caused errors consists of wholly repetitive variations such
as those caused by roll defects. Such errors are harmonic and the repeat period may
be expressed as error wavelength. A harmonic error produces a series of single
ordinates on a spectrogram, and the highest of these lines is usually the fundamental
wavelength. If a delivery roll is purely eccentric without any other error, the fundamental
error wavelength in the output strand is the circumference of the front roll. A roll with
a flat on it produces a fundamental wavelength and a number of harmonics (at 1/2,
1/3, 1/4, etc., of the fundamental wavelength). The defect is often caused by leaving
the load on the rolls when the machine is not running. Occurrences of these harmonics
in the test data should raise questions about the work practices in the mill. It is
relatively simple to estimate the source of each spike. The spectrum of mechanical
292
Handbook of yarn production
Roll separation
(arbitrary units)
peaks shown in Fig. 11.10(a) was made by recording the changing separation of the
centers of the front rolls of a drawframe [8]. No textile material was passing through
the frame at the time. The technique exposed the mechanical errors more precisely
than testing the sliver; furthermore, the use of an encoder made possible exact
correlations between the error and the result. The spikes at W, X, and Z were from the
top front roll, W and X representing the second and third harmonics, respectively.
(The spike Y came from another source and need not be discussed here.) On a
spectrogram of sliver from a drawframe, the mechanical errors also produce spikes
but it is unavoidable that there is some fiber-related error as well. A single spectrogram
is often not very useful in determining mechanical errors because of the profusion of
random spikes from non-mechanical sources. If spectrograms are taken at precise
time intervals and they are assembled as a three-dimensional graph, then if the spikes
persist throughout the time dimension of the graph then the cause is mechanical and
they represent a truly harmonic variation. Random spikes can be separated from the
harmonic ones in this way [2,9]. A faulty element in the drafting system causes
organized ‘spikes’ to appear, as shown in Fig. 11.10(b). This example is an especially
bad case, included merely to emphasize the difference between the two types of error.
Z
X
W
Y
1.66
2.5
5
Error wavelength (inches)
(a) Mechanical errors
Mechanical errors in spectrograms
Amplitude
Non-mechanical errors
m
Ti
e
Error wavelength
(b)
0.5 in
3 in
11 in
Fiber-related peaks
A
B
C
1 yd
Worsted
Carded cotton
Cut-staple MMF
Continuous
filament
(c) Fiber-related errors
Fig. 11.10
Three views of error spectra
10 yd
Quality and quality control
293
Fiber-related errors are more difficult to diagnose. The sources of error are many,
and they are dispersed. They range from (a) variations in the fiber selected, through
(b) the maintenance and setting of the machines, to (c) the environment in which the
operation occurs. A starting point is the theoretical random error that is dependent on
the number of fibers in the cross-section; consequently there are different values to
be expected from sliver, roving, or yarn. The evenness also depends on the fiber
length and its variability. Thus, there is a spectrum of values according to the strand
involved, as indicated in Fig. 11.11(a). A yarn contains errors from spinning, roving,
drawing, carding, and other processes. The error wavelength carried forward to the
yarn gets ever longer as the initial generating point is positioned nearer the beginning
of the process line. Also performance of one machine in a serial line of processes
Yarn
1
Short-term CV (%)
20
Roving
Sliver
2
10
4
3
2
4
5
1
2
3
4
=
=
=
=
Worsted
Carded cotton
1.5 den 1.5 in polyester
combed cotton
2
4
Theoretical
1
10
100
Linear density (tex)
(a)
CV of yarn mass (%)
25
95% USP
50% USP
20
15
5% USP
Carded
10
5
1000
Combed
USP = Position in the Uster statistics percentile (USP) rankings
= Results from a series of mills (combed yarns)
CV of yarn mass (%)
100
10
Linear density of ring yarns (tex)
(b)
20
Actual
15
Theoretical
10
1940
1960
1980
2000
Year
(c)
Fig. 11.11
Three views of strand variations (CV)
294
Handbook of yarn production
affects the following machines. A simple calculation can show that errors due to
carding produce yarn errors in the order of a million yards. Errors from sources
earlier in the process line produce even longer errors in the yarn. At the other end of
the scale, the delivery rolls of a cotton ring spinning frame produce errors of only 4
or 5 inches. Fortunately, the errors produced from the early stages are smaller than
those from the later ones. Mechanical errors from the ring frame produce short,
highly organized errors that produce moiré. Errors from the early process stages
produce barré in the fabric. The very long error wavelengths arising from the early
processes produce bobbins that contain yarns that vary in average count. The spinning
bobbins can become mixed between spinning, winding, and shipping. Consequently,
the fabric making equipment often has to deal with random step changes in yarn
count that show up as barré. Controlling the whole line is a technological art based
on experience as much as on science and technology.
11.3.3 Irregularity in linear density due to staple fiber variations and their
interactions with the machinery
Figure 11.11(a) shows some values of CV for various yarns and intermediate products
and it might be noted that an overall regression would not be parallel to the theoretical
line. Diagram (b) shows some carded ring yarn evenness values as judged on a
worldwide basis in the late 1990s. The values are for the error wavelength range 0.2
inch to 30 yd. Even the best yarns show CV values that range from 16% at 50s count
to about 10% at low counts. Longer-term variations, represented by the bobbin-tobobbin values, range from about 1% CV for the best spinners to a value for the worst
spinners that varies between about 3% and 5% at the counts mentioned above. These
long-term variations arise in the preparation.
Fiber-related errors in the last draft zone show up as distributions of variance on
a spectrogram that produce so called ‘hills’. They encompass a range of wavelengths
between one and ten fiber lengths, with the crest of the most common hill being
located between two and three fiber lengths. Wavelengths shorter than about half a
fiber length are unreliable and are not used. As with mechanical errors, the fiberrelated errors from previous drafting zones are elongated and the new crop of errors
is added. This means that there is a hill from the front draft zone, and sometimes a
less prominent one from the back zone. To get sufficient data to produce a spectrogram
requires the testing of a sufficiently long length of the strand concerned. This strand
then becomes waste fiber. Consequently it is not feasible to test for the errors from
the earlier processes as well as those from the current one with the spectrograph.
Materials from these earlier processes are tested immediately after the particular
process concerned.
Remember that the use of CVs does not discriminate between the sources of the
errors concerned, and an alternative method is needed. One way to minimize the
drafting errors is to (a) test the strand emerging from the subject machine for evenness,
(b) adjust the roll setting by one increment, (c) retest the strand and see whether or
not the amplitude of the crest of the hill has changed, (d) readjust the setting based
on the difference in amplitude, (e) retest, and so on until the minimum value is found.
By using the crests of the hills, one avoids the interference from other errors. The
‘hill’ discussed is not always very smooth and there are sometimes spikes that have
little to do with mechanical errors. One way of resolving this is to take a sequence of
spectrograms and array them. False spikes, unrelated to any mechanical error, will
Quality and quality control
295
stand out from the array, while the truly harmonic ones ‘march’ across the array in a
steady straight line. Experience over many years has made possible the production of
the Uster statistics, which give a good guide for judging performance. An example is
given earlier, in Fig. 11.10(b), where the shaded lines represent the different percentiles
of evenness from spinners throughout the world. Somewhat similar but more
comprehensive diagrams published by Uster Corporation are very useful for normalizing
data to compare plants on a meaningful basis.
Rather than refer to a particular CV at a certain count, it is better to refer to the
position in the Uster statistics percentile rankings (USP). For example, a mill spinning
30s then can be compared to another spinning 10s with reasonable results. Of course,
differences in fiber, machine maintenance or operational difficulties intervene, but a
normalized comparison is likely to expose difficulties that otherwise could be missed.
A group of mills can accumulate their own data and pursue a similar strategy. There
is a frequent practice of testing competitors’ products to get a basis of comparison but
care has to be taken to preserve an ethical stance on such testing. Figure 11.11(b)
shows some mill data of the early 1990s compared to the 1997 Uster statistics [1].
The particular individual mills belonged to a commercial group and their results then
fell in the 50 to 60 percentile range (the lower the percentile, the better). However,
mill data within any manufacturing group commonly vary less than with international
experience because of differences in the range of skills and equipment used.
Nevertheless, experience has shown it to be a powerful managerial tool because, once
norms are established for the group, logical comparisons can be made despite differences
in product and equipment. A similar procedure can be used for the Uster data regarding
thick and thin spots as well as neps.
Over the years, there has been steady improvement in the spinning equipment, the
methods of testing, and the application of the technology. The data given in the
foregoing are affected by this and some idea of the change can be obtained from Fig.
11.11(c). For example, the CV of mass for a 60s combed cotton yarn has declined
over a half century. Thus care has to be taken to keep the standards up to date; the
values steadily approach the minimum theoretical values. The same is true of other
fibers and preparations. This is one of the reasons why defects have assumed such
relative importance.
11.3.4 Irregularity of yarn hairiness
Hairiness of yarn is a factor in the appearance of fabrics, and variations in hairiness
can produce optical effects that show up as streaks or bars or other visual disturbances.
The Uster hairiness value, H, represents the total length of all the hairs protruding
from the yarn, in cm, with reference to a sensing length of 1 cm. It is measured using
infra red light to avoid color problems [9]. In an average carded cotton ring yarn on
the bobbin, H varies from about 10 for 6s (≈ 100 tex) yarn to about 4.5 for a 50s
(≈ 12 tex) yarn [1]. Subsequent winding and unwinding changes these data dependent
on the conditions prevailing at the two operations. The variation between the best and
the worst yarns is about 30%. The values quoted are averaged over a substantial
length of yarn and there are significant short-term variations. Uster report withinbobbin variations which are in a range that can be expressed as 18 to 28% CV. The
between-bobbin values vary between 1.2 and 10%.
296
Handbook of yarn production
11.3.5 Irregularity of yarn strength
Yarn strength for a staple yarn varies from about 10 cN/tex for the weakest yarns to
30 cN/tex for the best ones. A graph of the tenacities of various sorts of yarn is given
in Fig. 11.12 and it will be seen that rotor yarn has lower values than ring yarn. The
values for long-staple cottons for very fine yarns (not shown in Fig. 11.12) yield up
to 30 cN/tex tenacities. At the bottom end of the scale, poor worsted yarns have
tenacities around 7 cN/tex, poor rotor yarns produce about 10 cN/tex, and poor
combed 65/35 P/C ring yarns produce about 19 cN/tex. Tenacities appear higher
when tested at 400 m/min.
Figure 11.13 shows the CV of tenacity for various yarns. The best cotton yarns
vary from about 4% for 6s carded ring yarn to about 9% at 50s count (≈ 12 tex). The
worst yarns have CVs of the order of 10%. Combed cotton yarns have CVs varying
from 4.5% at 15s count (≈ 40 tex) to 12% at 120s count for the best yarns. The worst
yarns vary from 8.5% CV to 15% CV for corresponding counts (≈ 5 tex) and worsted
yarns have high CVs of tenacity. It should be noted that the CVs of tenacity measured
at higher testing speeds tend to be a little higher than those quoted.
11.3.6 Faults in fabrics
Coefficients of variation of the yarns do not tell the whole story; acceptability is
conditioned by the wavelengths at which the errors occur. Spectrograms, such as
Tenacity, log scale (cN/tex)
CV of tenacity, log scale (%)
Fig. 11.12
Combed 65/35 P/C, ring
30
Combed cotton, ring
Carded cotton, ring
20
0 P/C, air-jet
Carded 50/5
10
Carded cotton, rotor
Worsted
5
20
50
100
200
Yarn linear density, log scale (tex)
Comparison of yarn tenacities (best 5% of yarns tested at 5m /min)
20
Worsted
10
Carded cotton, rotor
50/50 P/C, rotor
50/50 P/C, air-jet
Carded cotton, ring
5
Combed cotton, ring
20
Fig. 11.13
Carded 50/50 P/C, rotor
Combed 65/35 P/C, ring
50
100
Yarn linear density, log scale (tex)
200
Average CVs of tenacity of some staple yarns (tested at 5 m/min)
Quality and quality control
297
shown earlier in Fig. 11.10(c), show that different fibers produce the majority of their
short-term errors at different wavelengths [9]. The area under the curves is a function
of the variance in the yarn; the pattern of peaks is as important as the magnitude of
the variation. The pattern affects what is seen in the fabric; the CVs of mass give only
an overall value, with no wavelength component [10].
Another way of illustrating the same point is to observe the fabrics. Knitted fabrics
are sensitive to yarn errors. A selection of photographs of knitted fabric appearances
and their accompanying spectrograms is shown in Fig. 11.14. In the space available,
only a taste of the subject can be given, but there is some value in discussing the types
of error shown. Picture (a) is of fabric made with normal yarn and serves as a
reference. Picture (b) shows the effect of poor fiber control in the front drafting zone
of a ring frame – the mottled look will be noted – and picture (c) has errors produced
in both front and rear drafting zones. The long waves from the rear zones produce the
wood-grain effect. All yarns were 20s cotton. Experience has shown that improvement
of short-term error without a corresponding improvement in long-term error leads to
the production of fabrics of unacceptable visual character.
(a)
(b)
(c)
Fig. 11.14
Fabric faults arising from yarn errors
298
Handbook of yarn production
11.4
End-breaks and quality
40
20
10
10
30
50
100
Yarn count, log scale (Ne)
Mean end-break duration (min)
End-breaks/1000 spindle hr
11.4.1 End-breaks in spinning
The topic of end-breaks has already been mentioned. From an operational point of
view, the end-breakage rate is a symptom of how well a plant is running. A high endbreakage rate points to a combination of machine, material, and human faults. If (a)
the travelers need changing, machine maintenance is in arrears, or the machine is
otherwise defective, or (b) the roving is bad, or there is a poor choice of fiber, or (c)
the operators are not performing well, then a high end-breakage rate will result. If the
training is poor, the assignment is too large, etc., then poor results should again be
expected. The design of the machine is also a factor and so is the use of monitoring
(human or otherwise). When all these factors are controlled, the author’s experience
suggests that the end-breakage rate is then quite a strong function of yarn count (Fig.
11.15(a)). Some authorities have different experience and their rate is less variable
with count because their range of expertise and equipment is much wider than in the
case quoted. The mean duration between end-breaks is shown in diagram (b); three
lines are shown and one might say that the lines represent the best, the average, and
the worst spinners in the world. As has been mentioned elsewhere, the end-break
duration is a matter of assignment and operator training. It also depends on the duties
assigned to the operator other than piecing-up. It is fairly obvious that the end-break
rate is a function of spindle speed but it is less obvious that it is a function of how
long the traveler has been in service. The rate increases similarly to the case shown
in diagram (c). Replacing travelers more frequently adds costs to production but it is
offset by the reduction in costs from lower end-breakage rates. The ends down rate
also is very strongly affected by the number of yarn defects/unit mass (usually compared
on the basis of defects/bobbin) as shown in Fig. 11.15(d).
60
40
90%
50%
20
10%
10
10
20
40
100
Yarn count, log scale (Ne)
(b)
15
Effect of traveler wear
10
5
0
200
400
600
Traveler use (hr)
(c)
Fig. 11.15
800
Ends down/100 spindle hr
Ends down/100 spindle hr
(a)
40
y = 10.8x + 11.33
30
20
10
0
1
2
Defect/bobbin
(d)
End-breakage in ring spinning
3
Quality and quality control
299
One inexpensive and effective way of detecting the places where excessive endbreaks occur is for the supervisor to tour the ring room looking for deformed bobbins.
When an end breaks, the building mechanism continues to work. The result is that,
when an end is repaired after a delay (which is normal), the package looks like that
shown in Fig. 11.16(b) instead of Fig. 11.16(a). The longer the delay, the greater is
the deformity. Greatly deformed bobbins mean that (a) the operator assignment is too
high, (b) the operator is poorly trained, or (c) the operator is inefficient. A high
frequency of widely scattered deformed bobbins means (a) the machine maintenance
is substandard, (b) the settings are wrong, (c) the sliver or roving supplied is substandard,
or (d) the operator assignment, training and efficiencies have to be reviewed. Such
tours should be frequent because it is only possible to make judgments on frames
where the bobbins are built sufficiently to make the deformities evident. Thus, a
single tour misses the frames where a doffing has occurred fairly recently. The tour
can also be used to check the spindles. If viewed with a simple flashlight, the balloon
can be seen easily. If a bobbin is out-of-plumb, it is quite possible that the upper tip
of the bobbin will touch the yarn in the balloon when the rail is in a low position.
Also, if a wrongly sized traveler is used, the same thing might happen. Such interference
causes intermittent hairiness in the yarn which, in turn, leads to hairiness, moiré, and
barré in the fabric. In addition, the use of too light a traveler can produce temporary
balloon collapses which are likely to cause end-breaks. The collapse can be seen with
the flashlight. The phenomena just described are seen most at the low rail position,
soon after doffing. A localized pattern of end-breaks is often found near doorways,
(a)
Fig. 11.16
(b)
Normal and deformed ring bobbins
300
Handbook of yarn production
local heat sources, or in a group of machines supplied with substandard sliver or
roving.
An accurate way of accomplishing the objective just discussed is to use a monitoring
system, but this is a somewhat expensive option. For ring spinning, one has to weigh
the advantages and disadvantages of monitoring. For more highly productive machines,
the advantages usually outweigh the disadvantages. Thus rotor spinning, air-jet spinning,
and many processes associated with the production of filament yarns are likely to be
fitted with a defect monitoring system.
11.4.2 Quality control and economics
As an example of the rate of production of faults, if production of a 24s yarn
is 1000 lb/hr and the spinning fault level is only 1 in 106 yards, then 24 × 840 ×
1000/106 = 20.16 spinning faults are still produced per hour. (See Q1 Appendix 2 for
the length of yarn in 1 pound.) If the fault rate is worse than this, the burden increases
proportionately. High fault rates cause losses in efficiency of spinning, winding,
twisting, beaming, and weaving. They also degrade the product. The degradation
increases the complaint level, reduces the price that can be demanded, and reduces
the volume of sales. As mentioned elsewhere, many of the spinning faults are caused
by poor preparation.
It is hoped that the foregoing illustrates that the decision of how and where the
yarn or the preparation should be sampled is not a simple one. To be economic and
effective, it is essential that the multitude of spindles in a ring spinning mill should
be properly monitored. Also, it is of similar importance to monitor the earlier stages
of production properly, particularly regarding those errors and faults that lead to later
trouble. Frequently, it is necessary to backtrack to find the cause of error. This
requires that the product be properly labeled. Testing is only effective when it is
accompanied by good organization and interpretation. Similar thoughts apply to the
production of textured and filament yarns. Many of the problems in the final product
have their origins in the early stages of yarn production.
References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
Färber, C and Furter, R. Uster Statistics No 40, Zellweger Uster, Uster, Switzerland, May
1997.
ASTM D2253 Measurement of Cotton Color with the Nickerson-Hunter Colorimeter, Textiles,
ASTM, Philadelphia, Annually.
Lord, P R. Unpublished data from the author’s private records.
Anderson, S L. Textile Fibres: Testing and Quality Control, Manual of Textile Technology,
Textile Institute, Manchester, UK, 1983.
Hattenschweiler, P and Bühler, M. Uster Bulletin No 21, Zellweger Uster, Uster, Switzerland,
Nov 1973.
Thomason, W A. A New Era in Quality Control – Yarn Fault Management, Uster News
Bulletin No 18, Uster Corp, Charlotte, NC, Jan 1971.
Douglas, K. The Uster Automatic Electronic Yarn Clearing Installation, Uster Bulletin No
22, Uster Corp, Charlotte, NC, USA, 1974.
Lord, P R and Grover, G. Roller drafting, Text Prog, 23, 4, 1993.
Douglas, K. Uster Bulletin No 35, Uster Corp, Charlotte, NC, USA, Oct 1988.
Douglas, K. Uster Bulletin No 15, Uster Corp, Charlotte, NC, USA, Jan 1971.
12
Economics of staple yarn production
12.1
Yarn economics
12.1.1 Cost and price
At the risk of stating the obvious, a distinction between cost and price must be drawn.
Costs are incurred by the producer and the price is what is asked upon selling the
product. Under non-monopoly conditions, price is determined by availability of the
product already in the market and the quality of the product being offered for sale.
The difference between price and cost is profit (or loss). The cost of a yarn may be
divided into several categories. More than half is from the cost of fiber, part represents
sales costs, and the rest comes from costs associated with the conversion of fiber (in
the bale form) to yarn (wound on cheeses or cones). Buying fiber is a very important
factor, not only in controlling costs, but also in determining mill performance. Buying
involves specialized skills that require an ability to assess value in the raw material.
A knowledge is required of what is needed to produce a yarn satisfactory to the
customer and what is needed to give acceptable performance in the mill.
12.1.2 Conversion costs
Conversion costs are a major component of the total. Evaluation of conversion costs
involves a knowledge, not only of how the fibers interact with the machinery, but also
of the commercial importance of yarn quality and cost. Major subdivisions of the
conversion costs are (a) direct labor costs, (b) overhead costs, (c) capital costs of
machines, (d) space costs, and (e) power costs. An assessment of cost proportions in
1995, calculated from ITMF data, indicated that, for a 30/1 combed cotton rotor yarn
produced in the USA, 52% of the costs were attributable to the cost of fiber. Some
conversion costs are shown in Table 12.1. The extra combing process (which is not
universally used) affects the issue. Nevertheless, it is clear that labor costs in rotor
spinning no longer dominate the conversion cost structure. This is different from
previous decades when the labor cost dominated. Despite this, the 1990s saw a surge
in ring spun cotton products for apparel coming from the Far East. In the USA, rotor
302
Handbook of yarn production
Table 12.1
Conversion costs for 30/1 combed cotton rotor yarn in 1995
Labor
Depreciation
Interest
Waste
Energy
Aux materials
10.4%
41.5%
18.5%
15.5%
8.3%
6.0%
spun cotton yarn has become common for denim products and ring spinning has
become increasingly confined to underwear products; most other apparel products
are imported and almost all of these are ring spun. A basis for estimating the current
yarn market is provided by the sampling used by Uster Corporation for their statistics
[1] and the result is shown in Table 12.2. Notable facts are that (a) the USA and
Europe are prime users of rotor spinning, (b) the USA now has a diminished ring
spinning capacity, (c) Asia is a strong supplier of ring spun yarns, and (d) Europe is
the primary manufacturer of worsted yarns.
Some private data are given in Table 12.3 and other private data show that the cost
of fiber in some more recent operations in advanced regions of the world can reach
70% of the total. Operating expenses in more recent operations are generally closely
guarded secrets and it would be improper to disclose them here. The relative reductions
of labor costs have a strong influence on the conditions of international trade but this
has been offset by the effect of international trade agreements. The above data may
be compared to the figures presented by Thompson [2] in 1982. He showed that in
North America the breakdown for a 30/1 yarn was 26.4% labor, 43.9% capital (i.e.
depreciation), 10.6% for energy, and 18.8% for space. These may be compared to
9.3%, 60.7%, 19.1%, and 10.9%, respectively, for the Far East. The total cost for Far
Estimated yarn supply percentages in 1997
Table 12.2
Cotton
Cotton
Cotton
65/35 P/C
50/50 P/C
Polyester
Worsted
Ring
Ring
Rotor
Ring
Rotor
Ring
Ring
Table 12.3
Labor
Overhead
Depreciation
Energy
Totals
Carded
Combed
Carded
Combed
Carded
W Europe North
Asia
America
South
Africa, E Europe Total
America Middle
East
29
47
35
3
25
38
70
25
7
11
8
–
11
15
3
3
25
3
68
11
–
15
29
13
39
–
25
–
24
10
12
34
2
4
8
4
4
4
13
5
11
7
100
100
100
100
100
100
100
Percentage conversion costs for 36/1 combed cotton ring yarn
Prep
1978
Spin
Wind
Prep
1988
Spin
Wind
43.3
26.1
15.2
15.4
100
52.9
21.9
10.9
14.3
100
52.2
22.4
13.4
11.9
100
37.0
23.9
23.6
15.4
100
39.0
21.7
18.6
20.6
100
43.8
27.1
18.7
10.3
100
Economics of staple yarn production
303
Eastern countries was 72% of the North American cost. European total costs were
said to be similar to the North American ones. The reason for the move of yarn
manufacturing centers to the Far East is quite clear, as are the inequalities in capital
investment, energy, and labor costs. It also helps explain the strong showing of rotor
spinning in the USA, where the increased investment has led to substantial labor cost
reductions, which have helped maintain viability.
In ring spinning, labor costs are higher. A private analysis (relating to a different
European area) yielded proportions similar to those given in Table 12.3. Accounting
procedures, the count and the proportions of labor cost/hr, power costs, etc., differed
in this case as compared to the data cited by Thompson. However, the higher labor
costs, which still dominate in the ring spinning of medium to fine yarns, are quite
apparent. In the decade illustrated, labor costs for the cases cited were reduced by
about 10% by investing in new and more productive machinery. The corollary to this
was the rise of more than 7% in depreciation costs which was associated with the
increased investment. Energy consumption in some mills has almost doubled over the
last ten years.
12.2
Productivity
12.2.1 Normalized productivity
One measure of mill efficiency is the number of operator hours needed for a given
task. This is a good measure because it enables comparison between mills running in
various environments; it is not affected by the wage rate and it can be expressed in
different ways. One way is to measure the number of operator hours to produce
100 lb of yarn. This form of normalization is represented by the acronym OHP. In the
metric system, the unit of mass is 100 kg and the applicable acronym is HOK; the
values for HOK are 2.2 times larger than for OHP. Basically, the units are expressed
in operator hours/100 lb or kg.
12.2.2 Historical changes in operator productivity
Technologies continue to change and enable mills to run with decreasing amounts of
labor. However, as the labor costs are reduced by applying new technology, the
competitive advantages between high and low labor cost regions diminish. Over the
last half century, the OHP level has decreased to less than 10% of what it was. This
is, in large part, because of the progressive introduction of various schemes of shortened
process lines and automation. Consider some practical data, which make the point.
Each year sees a development in technology that further reduces the values, as shown
in Fig. 12.1. The exponent of the regression curve (–0.074) suggests that the improvement
each year has averaged nearly 15%.1 The equation is exponential, which implies that
the rate of improvement will steadily decline as the labor use factor approaches some
steady value asymptotically. This estimate relates solely to one group and others may
differ considerably, but the trend is common wherever considerable investment has
been made. Suffice it to say that there is abundant evidence to show that labor
productivity varies with time, as equipment designs evolve.
1 10 –0.074 = 0.843, thus the value of OHP is, on average, 0.843 that of the previous year.
304
Handbook of yarn production
6
30/1 cotton
OHP
4
2
y = 6.454 × 10–0.074x
x = year – 1980
0
1980
1985
1990
1995
Year
Fig. 12.1 Historical decline in OHP
12.2.3 Division of operator productivity within ring spinning
The values of OHP or HOK can be broken down by components. In rotor spinning,
there are no separate winding costs as there are with ring spinning, and preparation
costs are lower because of the omission of the roving stage; however, finer slivers are
needed. Thus, labor costs are lower for rotor spinning than for ring spinning in the
viable range of counts (Fig. 12.2(a) shows some old data). In short-staple ring spinning
of fine yarns, preparation (all processes up to and including roving) takes up to 5%
of the total, winding takes up to 4%, as do the combined overhead costs of management,
maintenance, and general mill expenses. The rest is for spinning.
For yarns towards the coarse end of the count spectrum, preparation costs are a
much larger proportion of the total. Of that, a significant part is labor cost. When
expressed in terms of OHP, the example shown in Fig. 12.2(b) shows the labor needed
for spinning rises rapidly with count as compared to that needed for preparation and
winding.
Stryckman [3] and others have developed a series of HOK or OHP curves for
various fibers and spinning systems. The types of machinery in use, the preparation,
and the fiber lengths and types determine the equations. The data are good only for
the time at which they were produced and for the equipment then in use. Changes are
to be expected in future years. For fine counts of ring yarn, the OHP is almost directly
proportional to count, as shown in Fig. 12.2(c) (each point represents a single mill).
Regressions for the data are given. Data for other years have been plotted and, as
expected, the results suggest that the values are time dependent. The linear relationship
seems to hold up well for cotton counts above 36s. Linear regressions are given. In
the five year interval shown in Fig. 12.2(c), the OHP for a 40s combed yarn appears
to decrease from 2.27 to 1.66, which suggests a reduction of just over 5% per annum.
On a worldwide basis, the OHP levels off for coarse carded counts, and it is much
more variable from plant to plant. In the past, regions of lower wages produced less
efficiently than more advanced regions and, on a worldwide basis, performance was
represented as a fairly wide band rather than a single line. With wider markets and
competition, the width of that band seems to be narrowing. The balance of technology
has changed and many regions of lower wages now have modern machinery.
Some idea of the relative magnitudes and the relationships to count is shown in
Economics of staple yarn production
Preparation
Ring spinning
Winding
3
OHP
10
Ring
HOK, log scale
OHP, log scale
100
10
4
100
y = 3.646 × 100.010x Ring
y = 1.817 × 100.017x Rotor
305
2
1
Rotor
1990
1983
0
10
20
30
40
Yarn count (Ne)
(a)
5
50
10
20
30
40
50
Yarn count (Ne)
(b)
60
70
1985 y = 0.080x – 0.934
1990 y = 0.077x – 1.425
OHP
4
3
2
1
0
10
Fig. 12.2
Ring spinning
Each point represents 1 mill
20
30 40 50 60 70
Yarn count (Ne)
(c)
80
Normalized operator productivity
Fig. 12.2(b). It will be noted that spinning increasingly becomes dominant as the
count rises above 30s cotton. The balance between these is subject to changes as
automation is progressively brought to bear on the labor costs in those technical
areas.
For coarse count yarns, materials handling costs, which account for much of the
labor charges in preparation, assume great importance. It is also interesting that
winding costs become progressively more important at the lower counts. At one end
of the count spectrum, quality and piecing costs are of great concern, whereas at the
other end, materials handling becomes a dominant problem.
12.2.4 Operator productivity in rotor spinning
With modern industrial rotor spinning, the labor costs are about 50% of those for ring
spinning. A rotor machine is often fed with sliver produced by an abbreviated preparation
line; sometimes only one passage of drawing intervenes between the card and the
rotor spinning machines. This reduces costs. With modern rotor spinning machines,
automatic piecing is a necessity because human operators are not capable of carrying
out the operation reliably at the rotor speeds now possible. As with ring spinning,
automatic doffing is standard. Consequently, jobs that were once performed by human
306
Handbook of yarn production
operators are now done by machine. Also, the productivity of the rotor machines on
a count-for-count basis is over five times that of a ring frame. Obviously rotor
spinning falls into a different category, as do most of more advanced forms of spinning.
Typical values of the ratios of OHP/count are 0.025 for rotor spinning, as compared
to 0.05 for ring spinning.
12.2.5 Operator assignment
Another way of judging mill efficiency is by the operator assignments. An operator
of a ring spinning machine might tend many hundreds of spindles [4] and that would
be his/her assignment. Assignments have increased appreciably over the years. There
are some interesting consequences arising from the large number of items being
controlled by a single person. Inevitably, some ends break whilst the operator is
patrolling the spindle set. The cost of this operator has to be set against the savings
he or she can make by repairing the broken ends. The cost of the operator depends on
the wage rate but the assignment depends on the economic balance at which minimum
cost, or acceptable quality, is achieved. The literature over the last half century was
searched to find how the assignments and wages had changed and the results are
plotted in Fig. 12.3(a); regression curves shows the differences between two common
yarn types. Differences are to be expected because cotton is a natural fiber and more
variable than polyester. Again, care has to be taken with comparisons, especially
since the operator often has duties other than piecing.
12.3
Quality and economics
12.3.1 Yarn quality and operator assignment
Figure 12.3(a) compares spindle assignments across the world. One can note that, in
1980, the higher the wage paid, the higher was the spindle assignment. Equation
[12.1] on page 308 shows that this should be expected. However, the equation can
take into account neither the range of technologies in use nor the influence of differing
qualities of product from the various systems.
Poor yarn quality not only degrades the price that the product will fetch, but it also
imposes a cost penalty in spinning and winding. Logic suggests that operator assignment
should also be influenced by the quality of the yarn produced [5]. The more weak
places in the yarn, the greater is the end-breakage rate and the higher is the need for
labor. Plotting data from a set of mills against their spindle assignments for like
products (Fig. 12.3(b)) shows that the spindle assignment drops tremendously when
the number of thin spots rises. The regression shows that the assignment for these
cases was just about inversely proportionate to the number of thin spots per unit
length. Some quality factors affect the immediate cost of production but the effects
of others are not seen until a later process, possibly in the plant of a customer. Let
these be called Category A and Category B problems, respectively. The latter are
more difficult to deal with because the operators and their supervisors do not feel the
effects at first hand. Nevertheless, they are very important. The major effect of
Category B problems is to undermine the price of the product. Category A problems
reduce the efficiency of operation but do not undermine customer confidence unless
they are allowed to continue uncorrected.
Economics of staple yarn production
Spinning assignment
(spindles/operator)
Spinning assignment
(spindles/operator)
10 000
40s Polyester/cotton
1000
20s Cotton
100
0.1
1
Wage rate (1980 $/hr)
(a)
ye =
Breaks
1000 spindle hr
100
1 000
100
10
10
Break duration
y = 25 777 x–0.96
x = thins/km
100
Thins/km
(b)
yd = 2.760x0.601
Break rate ye = 2.693x0.579
x = yarn count (Ne)
100
Duration
Break rate
10
10
All scales are logarithmic
Yarn count (Ne)
(c)
10
100
1000
yd = break duration (min)
y = 772x 0.92
y = 410x 0.56
x = wage rate ($/ hr)
10 000
307
Fig. 12.3 Factors in determining spinning assignment
12.3.2 End-breaks and operator assignment
Operator assignment is strongly affected by the end-breakage rate in processing.
Consequently, no discussion of the economics of spinning should avoid this aspect.
A set of performance figures for a number of mills is shown in Fig. 12.3(c). As the
number of ‘thins’ increases, so does the number of weak spots, and an operator can
only serve a smaller set of spindles. In other words the assignment has to be lowered.
Accurate data is difficult to acquire for several reasons, three of which are given.
First, methods used by managers differ from one mill to another. Second, different
fibers are in use. Third, there is a tendency for mill personnel to try to cast the
performance of their mills in a favorable light. The correlations are poor but clearly
both the end-break rate and the waiting time vary with yarn count. Other problems,
such as the number of weak spots, are also associated with count (and other factors).
Nevertheless, it is a reasonable expectation that fine yarns will be more difficult to
spin than coarser ones. This is one of the reasons why a more careful choice of fibers
should be made for fine yarns.
The product of the end-breakage rate and the duration of the down time enables an
estimate to be made of the lost production. Taking the data in Fig. 12.3(c) at face
value, the loss in production during the time a break waits for repair is yd minutes (or
{yd/60} hrs), the production between breaks is 1000/ye hrs and the percentage loss in
production is approximately yeyd/600. The maximum probable value of loss due to
end-breakage in those data was 2.37% and the loss was very roughly proportional to
count. On this basis, spinning a 30s count might bring a loss of machine productivity
of just less than 1%. Besides machine productivity losses, the cost of labor increases
308
Handbook of yarn production
due to the end-breakages. However, since the figures are very dependent on the fibers
and practices used in a given mill, they can do little more than give an order of
magnitude.
Correlation between the loss in production and pneumafil waste collection is
usually poor because pneumafil continues to be collected even when spinning continues.
Examining thousands of spindles has shown that fiber is always being removed from
the fiber stream entering the twist triangle during the spinning process. Furthermore,
the amount removed is affected by the quality of the fibers. The waste figure increases
if there is a large short-fiber content. Unfortunately, the short-fiber content is highly
variable at this point. Of the 2% or 3% pneumafil waste levels common in ring
spinning of cotton yarns, up to 0.6% is due to the continuous loss.
12.4
Cost minimization
12.4.1 Theoretical assignment
For simplicity, let it be assumed that X spindles in a set are equally spaced around a
circle and that end-breakages occur randomly. Further, assume that the operator deals
with each end-break in sequence and never turns back. Under stable conditions, as
the operator deals with one break and moves to the next, there is a probability that an
end breaks somewhere else in the spindle set; that spindle then becomes unproductive.
There will always be u spindles not producing, and, on average, they have to wait t
hours before they are repaired. During the waiting time, the set fails to produce ptu
out of the ptX lb/hr expected, where p is the productivity of one spindle in lb/hr. If the
assignment is increased, there is a saving in labor costs, but the efficiency of the
machine is reduced. There is a deterioration in the quality of the product because of
the multiple piecings, which have to be removed in winding. The assignment has to
be optimized and perhaps negotiated with the unions. It has been shown [4] that when
piecing dominates the work load, the optimum assignment, a, is approximately given
by:
a = √ [Cl/{Bt (pCw + Cf)}]
where
a
B
t
p
Cw
Cl
Cf
Ch
OHP
=
=
=
=
=
=
=
=
=
[12.1]
spinners assignment in spindles/operator
end-breakage (rate/hr)
average time (hr) for the operator to pass from one spindle to the next
productivity for one spindle (lb/hr)
cost of reprocessing the waste ($/lb)
cost of labor ($/hr)
fixed costs ($/hr)
handling costs ($/hr)
number of operator hours needed to spin 100 lb of yarn.
Bt is a function of yarn count, as is p. Thus the spindle assignment is affected not only
by the economic factors, but also by the count and quality of the yarn being processed.
Equation [12.1] indicates that the fixed cost is an important factor and the other
major components are the end-breakage rates, the size of the operator sets, and the
cost of labor. The cost of labor may be divided into three categories, namely (a)
supervisory and maintenance staff, (b) mill personnel dealing with materials handling,
and (c) personnel whose task it is to keep the machinery productive – this entails the
Economics of staple yarn production
309
repair of end-breaks as soon as possible after they break to keep up the operational
efficiency, as well as the regular maintenance. Category (a) is not normally considered
part of the direct labor force because service usually continues over long periods.
Consequently the costs from this source can be considered part of the fixed cost and,
for the present purpose, can be lumped with costs of servicing the investment and
direct maintenance costs. Category (b) is regarded as part of the handling costs, and
category (c) is part of the variable cost. Spinning operator assignment can also be
expressed in terms of OHP and machine productivity.
a = 100/[OHP × p] spindles/operator
[12.2]
12.4.2 Capital and fixed costs
Studies in 1990 covering many mills showed that the value of the machinery installed
in a mill rises almost linearly with yarn count. Before 1990, the cost was estimated
by various people to be the equivalent of up to $1 million per ton/day for each unit
of count. In the twenty-first century, a new mill producing 30/1 yarn might be expected
to cost perhaps $10 million for each ton/day capacity. In other words, the textile
spinning industry has become a capital intensive industry.
Despite the improvement in spinning technology, piecing a ring frame is still a
major consumer of labor. Piecing costs also rise roughly linearly with count (with the
factor standing at about 0.45¢/lb for each unit of count). Obviously the figures alter
with time, as changes in the design of the machinery modify the costs and performance;
also as currencies suffer inflation. Nevertheless, it is easy to see why the average
count produced tends to have reduced over the last few decades. However, there is
still room for the high count producer of specialty yarns, despite the cost.
In ring spinning, the productivity per spindle is very low and since expenditures
have to be paid for from the proceeds of sales, the amount that can be spent is also
low. For example, consider a spindle producing 0.02 lb/hr for 7000 hr/year in an
environment where the most extra charge that the market can bear is, say, 3¢/lb. The
most that could be afforded is little over $4 per year per spindle. Even with a ten year
payback, the most that could be spent is $40 per spindle! For a mill with 50 000
spindles, this latter figure is equivalent to $2 million, which is not a negligible sum!
All of this has to be taken into account when considering monitoring and computer
control. To the extent to which they can pay for themselves, they are fine, but there
is little margin for cost overruns. Of course, these cost estimates are transitory and
will change over the years.
In the early days, the capital (or fixed) cost of the rotor spinning machines was
so high that there was a break-even count above which the spinning system was
uneconomic. It was not until rotor speeds could be raised to improve the output per
dollar invested that the bar of a break-even count was overcome. With rotor speeds up
to 130 000 r/min now possible, the cost of building in features such as automatic
piecing can be absorbed into the original price of the machine without killing the
market. Early fears of substandard yarn have given way to acceptance over a wide
spectrum of products. Thus, we now have a category of high capital cost machines
that need relatively small amounts of labor to operate. This is in contrast to ring
spinning which has a relatively low capital cost but which needs more personnel.
310
Handbook of yarn production
12.4.3 Variable costs
The cost of labor is mostly viewed as a variable cost. However, the charges for certain
management and maintenance staff are regarded as administrative costs, which is
part of the fixed cost category (their costs are not directly linked to production and
therefore they are not viewed as direct labor costs). Of the variable costs, piecing of
ring frames is the dominant portion for fine-count ring yarn production. As previously
mentioned, piecing is automatic in rotor spinning, the end-breakage rate is less, and
these are two of the reasons why rotor spinning needs less labor. Variable costs are
important not only in the assignment equation but also as a measure of the mill’s
performance.
Automatic piecing for ring frames has been offered for sale by several machine
manufacturers. They were technologically sound but were not commercially acceptable.
The extra capital cost was an important factor in the lack of success. Nevertheless, a
lower cost solution might still be offered and this would change the theoretical model
that will be proposed later in this chapter. However, until an acceptable device becomes
available, it is useful to analyze the operation as it now exists.
12.4.4 Handling costs
In the first half of the twentieth century, the opening line consisted of manually
operated feeders and a number of cleaning and opening machines not physically
connected. Operators were used to transport material from one machine to another
and were also needed to feed the material into the system. In a modern plant, no
operator is required for material transfers in this section of a short-staple plant. The
system is now automatic from the time the bales are laid into position until the sliver
emerges from the card. The only personnel required in this department are supervisory,
except for the operator who puts the bales in the laydowns about once per day.
Between carding and spinning, operators are often required to move cans of sliver
from one machine to another, and a few operators are needed to supervise the machinery.
The introduction of automatic can changers and movers provides an alternative to the
job of sliver can moving but, again, it is a case of capital expenditure being balanced
against that of human operators. There is a similar choice in the zone between the
processes of roving and spinning. Several degrees of automation are available for the
transport and sorting of bobbins from the ring frame to the winder. On average,
considerable labor is still used today in moving sliver cans, roving bobbins, spinning
bobbins, and cones or cheeses of yarn. Bobbin transport systems and automated
handling of the final packages have reduced the handling costs for those who can
afford the capital outlay. By introducing automatic handling equipment, there is a
transfer from the handling cost category to the fixed cost one. It is these sorts of
transformations that increase the capital cost of the equipment and drive the industry
to become ever more capital intensive.
12.4.5 Cost proportions
Costs may be divided into categories such as are shown in Fig. 12.4. Some alter
almost immediately according to demand; these are described as short-term variables.
Costs in other categories change little or are not controllable by management and
they take place over longer periods of time; these are designated long-term variables.
Administrative costs vary according to the company structure and may include items
Economics of staple yarn production
311
Total costs
Conversion costs
Short-term
variability
Fig. 12.4
Profit
Overheads
and shipping
Rent
Depreciation
and interest
Fixed costs
Administrative costs
Power
Direct labor
Other materials
needed in processing
Fibers and/or filaments
Variable costs
Taxes
Material
costs
Long-term variability
Distribution of costs
not otherwise included in the foregoing. For the present purpose they will be taken as
long-term items. Investment in new machinery nearly always incurs higher charges
for depreciation and interest than the corresponding ones for the machines replaced.
In fact, the machines replaced have often been written off. These charges are placed
in the fixed cost category. Direct labor and power costs cover the variable costs in the
mill; these have been further divided into materials handling (for short, ‘handling’)
and other variable costs. If the total conversion cost is described by the three cost
components just discussed, then:
Total conversion cost = Cv + Ch + Cf
[12.3]
The variable cost, Cv, rises with spindle speed because of the increased end-breakage
rates and power consumption. For a given set of equipment, this cost also rises with
count because of (a) the diminished machine productivity at the higher count, (b) the
increased cost of energy, and (c) the increased end-breakage rates at high counts. For
simplicity, assume that the variable costs can be lumped together and described by Cv
= KN m, where K is a constant, N is the count and m is fixed. Since a log graph of the
lumped costs on the simplified basis is a straight line of slope m, it is useful to
express the relationship as:
Log Cv = Log K + m Log N
The number of yarn packages produced is proportional to N – 3/2 and the handling
costs, Ch, are proportional to those for a given set of machinery running at a specified
reference speed. By arguments similar to those used for the variable costs, the log
(handling costs) can be plotted as a straight line on a graph.
Log Ch = Log K′ – (3/2) Log N
[12.4]
Fixed costs, Cf, are, by definition, fixed. The calculation assumes that all the equipment
is in full production and there is no idleness due to malbalance of the mill. As will be
discussed later, theoretical balance in a mill with a single product is achievable only
in a plant that runs a single count. Where there is more than one product, it is difficult
to even approach perfect balance. Most mills have a degree of malbalance.
Figure 12.5 is a plot combining all three elements for each of two cases of differing
variable costs. A curve of total cost is given in each case. It will be observed that the
312
Handbook of yarn production
Total
Log (cost / hr)
C1
Cv 2
C2
Variable
C v1
Fixed
Handling
N1 N2
(a)
Log (cost / hr)
T1
Log N
T2
C2
Cf 2
C1
Cf 1
Ch1
Cv
Ch 2
N2 N1
(b)
Fig. 12.5
Log N
Optimizing costs
addition of the three cost components, to give the total costs, results in curves with
distinct minima. The symbol N is used for count and the appropriate subscript can be
applied to whatever count system is in use. The graph shows two curves for variable
costs Cv1 and Cv2, which could be caused, for example, by buying inferior fiber for
one case. At the count for minimum cost, the use of poor fiber increases the labor cost
from C2 to C1 and reduces the best count from N2 to N1. Thus the market becomes
more restricted because of the lower count. The competitive position will be eroded
if the additional cost shown to the right of N1 is not offset by the savings in fiber
costs.
For a given machine, the position of the minimum cost along the count axis is
controlled by factors other than fixed costs (Fig. 12.5(a)). (As an aside, if one were
to imagine the case of changing one rotor spinning machine for another, the fixed
costs/lb are increased but that would not affect the optimum count). Increases in
handling cost move the optimum count towards the higher counts and increase the
value of the optimum cost. Increases in variable costs move the optimum towards the
lower counts and increase the optimum costs. It is a matter of whether the handling
or variable costs predominate that determines the result. Remember that variable
costs comprise direct labor and power costs. Costs of labor and power can vary quite
widely from area to area, and this plays a significant role in deciding whether new
major investment is justified. For example, such an analysis might be used to see
whether rotor spinning should be brought in to replace ring spinning.
Economics of staple yarn production
12.5
313
Operational factors
12.5.1 Automation
Perhaps the most important step in automation of modern times was the successful
introduction of autodoffing. A similar process for the roving frame has been more
recently introduced but is not universally accepted. Also, an automatic device for
changing the travelers has been developed and this seems to work well. Each step of
automation decreases the handling costs in return for an increase in capital or fixed
costs.
Figure 12.5(b) shows the effect of changes in handling costs from Ch1 to Ch2. An
increase in fixed cost is inevitable and has to be taken into account as also shown in
the same diagram. The value has been changed from Cf 1 to C f 2. An arbitrary ratio
between the savings in labor cost and increase in fixed cost has been used for
demonstration. The effects of upgrading the automatic handling equipment are to
alter the best count (e.g. a shift from N1 to N2, as depicted in Fig. 12.5(b)) and the
optimum cost. Whether or not the total cost is improved depends on the additional
capital cost involved. The net result is that the total cost curve is flattened and the best
count might be lowered. Automation tends to be more favorable for heavier counts
and it is less sensitive to moves away from the best count. In a ring frame, automatic
doffing has become standard and is now often included in the original cost of the
machine. In rotor spinning and other forms of new technology, automation is almost
universal.
Automatic equipment leads to a substantial reduction in operator hours needed to
produce yarn but, as was said earlier, there is an extra capital cost involved. All these
schemes are technically feasible, but the problem is to finance them. The whole
process of substitution of capital for labor requires access to capital. Furthermore, the
final cost/lb must be no more than that of any competing systems, unless there is a
special feature given to the textile product that enhances its value. This is a major
reason for the slow adoption of the new technologies. On the other hand, if yarn
manufacturers who have to pay high wage rates are to compete, the labor content has
to be reduced and there will always be an interest in increased efficiency.
12.5.2 Mill balance
Unless a mill is balanced, some machines will lie idle or work at a reduced throughput.
Such a situation has an adverse effect on the economics of production. Ideally, each
process stage will process approximately the same flow rate of material. The productivity
of the spinning machine is a significant function of the count and is given by:
P = Uη/(504 TM Ne √N e )
where P
U
η
TM
Ne
=
=
=
=
=
[12.5]
machine productivity (lb/spindle hour)
spindle speed (r/min)
efficiency per spinning unit
twist multiple of yarn
yarn count (hanks/lb).
For practical purposes, all factors other than count vary little; the spindle productivity
is inversely proportional to N3/2 for a given spindle speed, since TM varies but little.
The throughput of the opening lines is little affected by yarn count and intervening
machines are weakly affected. Since mass flow must be conserved:
314
Handbook of yarn production
spinning output = output from prior stages of processing – waste generated [12.6]
A change in spindle speed, twist multiple, or spinning efficiency affects the
productivity mentioned in Equation [12.5] although the major factor is the count.
Hence, if there are m spindles in the mill per opening line, the output can be written
as PT ≈ mKN –3/2. If the net output of the opening line is Ko, then Ko ≈ mb KN – 3/2 at
balance, where mb is the number of spindles needed for balance, and mb ≈ (Ko /K)
N 3/2. However, Ko /K is a constant under the stated conditions and so mb is proportional
to N3/2. The balance is heavily dependent on count. In practice, a mill is never
completely in balance but with a stable order book, machinery is installed as necessary
and a rough balance may be achieved.
50
0
5
10
CV of yarn strength (%)
(a)
Statistical frequency
Ends down
1000 spindle hr
12.5.3 End-breakages and economics
The foregoing makes clear that spinning contributes a major portion to the cost of
yarn. In ring spinning, much of that cost is caused by end-breakage. In rotor spinning
and other new developments, this may be less so. In any case, it is worth considering
the aspect of end-breakage in yarn production.
One factor in the end-breakage rate is the CV of strand strength. The word strand
is taken to include the weak point in the twist triangle in ring spinning. An analysis
of cotton yarns from a variety of mills showed a relationship between the ends-down
rate and the CV of yarn strength (Fig. 12.6(a)). The correlation is imperfect because
of the varying conditions in the various mills but there is a definite trend.2 As the
yarn tension increases, the stress in the strand increases and so does the end-breakage
rate. The probability of an end-break depends on the probability distributions of
applied stress and tenacity of the strand (Fig. 12.6(b)). The area under the intersection
of the two distribution curves gives the probability of a break. The applicable tenacity
is that of the weak spots in the flowing material. In the case of ring spinning, this is
often located in the twist triangle and is a variable with time. In rotor spinning, the
weak spot is at the point where the yarn is removed from the rotor groove (except at
very high speeds). Under high speed conditions, surges in false twist increase yarn
tension temporarily at the navel and the yarn breaks occur there. There is an incentive
to reduce the yarn tensions and CVs of yarn strength. It is assumed that the count and
Applied stress
Grayed area =
probability of
an end-break
Yarn tenacity
Stress or tenacity in compatible units
(b)
Fig. 12.6 CV of yarn strength and end-breaks
2
One obvious factor ignored in the above is the spindle speed, and another is traveler weight; in
fact there was neither control nor measurement of the yarn tension which is a major reason for
the poor correlation. Unfortunately, adequately controlled data of this sort is very difficult to
gather in normal industrial practice.
Economics of staple yarn production
315
twist are set by the market and are not a variable. As far as the economics are
concerned, it is the number of piecings made and the number of bobbins containing
piecings that are important. One has a large impact on the cost of spinning; the other
impacts on the cost of winding and the quality of the product. There are both cost and
price implications.
To emphasize the last point, consider Fig. 12.7 in which the average number of
ring bobbins spun per end-break was calculated for a range of mills spinning cotton
yarns of various counts. Again, this was not a fully controlled experiment because it
is almost impossible to get data on a wide enough scale under fully controlled conditions.
Nevertheless, there is a clear trend and the correlation coefficient was better than
might be expected. The technical difficulties in spinning high count are clearly seen.
Quality control from a purely economic standpoint becomes ever more important as
the count rises. The number of bobbins with an end-break leaving the ring frame is
one important factor in determining the work load for the winder. If the end-breakage
rate in spinning is high, there is probably a high rate of intolerable yarn faults. Not
only will the winder have to remove most of the piecings but it will also have to
remove the accompanying yarn faults. The chance of a piecing in a bobbin is low at
fairly low yarn counts. On the other hand, the chance of an end-break within a bobbin
of high count yarn is much higher. The number of breaks per cheese or cone is, of
course, much higher still. There is a piecing for every ring bobbin used in making a
cheese or cone. In addition, there is a piecing for nearly every objectionable fault,
many of which are caused by end-breaks. A few faults escape through the aspiration
device used to thread up the winder. For example, at 70s count, a break in every other
bobbin might give 30 + 30 = 60 or so piecings per wound cone. At 30s, the figure
might drop to, say, 30 + 2 = 32 piecings per wound cone. As an economic matter,
clearly poor spinning begets a poor winder performance. Both undermine what otherwise
might have been good economic results. Perhaps the operator costs for piecing should
be based on the production they save by intervention rather than on the number of
spindles they serve.
12.6
International competition
Bobbins spun/end-break
International competition is complex and in the space available it is only possible to
outline some factors. First, let us make a comparison between a high cost, highly
automated mill in one region, and a less sophisticated mill in a region of lower wage
15
y = 0.012x 2 – 1.544x + 50.368
r 2 = 0.950
r = correlation coefficient
10
5
0
30
Fig. 12.7
40
50
60
Yarn count (Ne)
End-breakages per bobbin
70
316
Handbook of yarn production
Area of advantage for Case A
Cost/hr, log scale
Area of advantage for Case B
Case B
Variable
costs
Case A
Handling Fixed
costs
costs
Yarn count, log scale (Ne)
Fig. 12.8
International competition
rates (Fig. 12.8). The scenario in this case is that a mill in an advanced region with
high labor costs seeks to offset the disadvantage by using automatic equipment. This
means that the fixed costs are high in this case. The mill in a region of lower wage
costs feels less need to invest in expensive automation. There are trade-offs that
might be in favor of one group or the other, depending on costs. Remember also, that
relative conditions undergo continuous change and no fixed recipe can be recommended.
In the particular case shown, the highly automated, high cost mill shows an advantage
only over a certain low count range. The high fixed costs have the effect of leveling
the total cost curve so that it is not so sensitive to count changes. The spin limit forms
a natural boundary to the right of the diagram.
As wage rates tend to equalize, these distinctions tend to disappear, but equalization
of wages is a prospect for the distant future. A factor not included in the analysis is
that of shipping costs. Transoceanic shipping decreases the margin of profitability.
Since low labor cost producers are usually distant from the major markets in advanced
countries, shipping costs offset their advantage of lower labor costs. This could be
taken into account by adding shipping costs to the normal fixed costs. It might also
be noted that interest rates available in some regions are less favorable than those
granted elsewhere. This, too, influences the final cost. Further factors are those of
quality, reliability, and promptness of delivery.
References
1.
2.
3.
4.
5.
Färber, C and Furter, R. Uster Statistics No 40, Zellweger Uster, Uster, Switzerland, May
1997.
Thompson, A. Techno-economic Aspects of Textile-machinery Investment, Textile Machinery:
Investing for the Future, Textile Inst Ann Conf, 1982.
Stryckman, J. Une Méthode de Mesure de la Productivité du Travail et du Matériel en Filature
de Coton, Centexbel, Belgium, May 1983.
Lord, P R and Mohamad, S B. Economics, Technology and Development in Staple-yarn
Manufacture, Managing Technological Change, 64th Ann Conf, Textile Institute, Oct 1980.
Garde, E. Process Control in Cotton Spinning, ATIRA, Ahmedabad, India, 1974.
Appendix 1
Calculations I: Elementary theory
A1.1
Yarn and strand numbering systems
A1.1.1 The basic philosophies
Textiles are often sold on a weight basis and consequently it is natural to express the
fineness of a yarn in terms of mass (or weight). There are two basic ways in which
this may be done. These are: (a) by specifying how much a given length of yarn
weighs; or (b) by specifying what length of yarn there is in a given weight. These are
known generally as the direct and indirect systems of yarn numbering.
Direct number = mass/length
[A1.1]
Indirect number = length/mass
[A1.2]
It will be noted that one is the inverse of the other. In the first case, the number gets
larger as the yarn gets heavier, and in the second case it gets smaller. The term mass
has been used because this is technically correct, even though the popular term is
weight. Mass is the amount of material in an object and weight is the force that acts
on it when it is accelerated. Since we all live in a gravitational acceleration of about
32 ft/sec2, all objects are subject to a force acting towards the center of the earth; that
force is what we call weight. The same mass on the moon would weigh a different amount.
Each system has its advantages and disadvantages; each has found areas in which
it has endured and has become established by custom. It so happens that, because the
lengths are so very long for any reasonable mass, the yarn numbers would get impossibly
small or large unless special counting systems are used. (Within the normal range of
linear densities, one pound of yarn laid in a line would extend many miles. The
numbers are unwieldy.) The following paragraphs explain a selection of the most
important counting systems.
A1.1.2 Direct systems
The technical name for fineness is linear density1 and it is always expressed as
1 This should not be confused with the term density as used in physics.
318
Appendix 1
mass/unit length. In commerce, the technical name is used less than in the fiber
industry or scientific community, such units of measurement as denier or tex being
often used instead. Sometimes the term yarn number (explained shortly) is used, but
this can be ambiguous.
Two of the major subsections of the direct system will be cited. In one, the logically
minded scientists have chosen the metric system and use g/km (the unit is called a
tex). In the other, the technologists have chosen g/9 km (the unit is called a denier);
this is based on an ancient measure of length but it still survives because it happens
to be about the right size of unit to describe a typical fine fiber. The normal metric
prefixes can be used in the tex system. For example, a decitex is one-tenth of a tex
and a kilotex is 1000 times larger than a tex. One denier is equivalent to 1/9 tex or 10/
9 decitex.
The denier is a popular unit in the fiber industry and many fibers of 1.5 denier are
supplied to blend with cotton; in the tex system, the commercial equivalent is 1.5
decitex. (The 10% difference is normally ignored.) Microfibers fall into the range of
1 denier or less. A carpet fiber often runs at 15 denier. In passing, it should be noted
that a 450 denier yarn made up of 1.5 denier filaments would contain 450/1.5 = 300
filaments in the cross-section.
There are also intermediate products, such as sliver, to which the direct system
of measurements is applied. Sliver is a rope-like strand that is much heavier than
yarn; a normal linear density is about 5 kilotex. In much of the industry, a system
using grains and yards is used for sliver; typical values are in the range 30 to 100
grains/yd. There are 7000 grains per lb. In this book, the symbol n is used for
measurements in the direct system of yarn numbering and the capital letter N is used
for the indirect system. In all cases, the appropriate units of measurement should be
placed after the quantity.
A1.1.3 Indirect systems
Indirect systems utilize terms of length per unit mass. There is a large variety of
systems, which is a legacy of the ancient crafts. Generally all the systems in this
category are called yarn count or yarn number. The term yarn count is preferred. It
is normal to specify the yarn count in hanks/lb where a hank contains a specified
length of yarn; unfortunately each of the systems specifies a different length. Therefore,
it is helpful to always specify the sort of hank being used when quoting a yarn count.
Some specified hank lengths are listed in Table A1.1 With the indirect system, the
number gets larger as the yarn gets finer. In the English cotton system, a 4s yarn is
very coarse whereas a 50s yarn is fine.
In cotton processing (and those technologies that have evolved from it), the units
developed in England in the industrial revolution are still in use. A cotton hank is
defined as 840 yd of yarn. (The number 840 is divisible by 1, 2, 3, 4, 5, 6, 7, and 8;
one can imagine the value of that in early primitive societies.) Thus, if the count of
a singles yarn is 20 cotton hanks/lb, there are 20 × 840 yd in a pound of yarn. It
should be noted that the yarn count is usually written as 20s or 20/1. The symbol used
in this book is Ne, where the subscript refers to ‘English’ and distinguishes it from
Nm, which refers to the metric count (meters/gram). In the case of long-staple yarns,
where the technology is derived from one of the processes for making yarn from
wool, a worsted hank is defined as 560 yd of yarn. In this case the symbol Nw is used.
Other systems of symbols are used by others. The American Society for Testing
Calculations I: Elementary theory
Table A1.1
319
Strand numbering systems
Direct
Fiber
Cotton
Wool
Man-made
Denier g/9 km
Intermediate
Card Fleece
Kilotex
Roving (cotton)
Roving (wool)
Roping
Yarns
Man-made
All yarns
Cotton type
Worsted count
Woolen count
All yarns (European)
Indirect
Name
Units
Name
Units
micronaire
fineness
decitex
–
Approx µg/incha
mg/cm
dg/km
–
–
–
–
–
–
–
lap weight
sliver weight
g/m
tex
–
–
oz/yd
grains/yd
–
g/km
–
–
–
–
–
hank roving
roving weight
roping weight
–
–
denier
tex
–
–
–
Metric count
g/9 km
g/km
–
–
–
m/g
–
–
–
–
English cotton count ch/lbb
Worsted count
wr/lbc
Woolen count
wo/lbd
Metric count
m/g
ch/lbb
wr/lbc
wo/lbd
Notes: (a) This is an arbitrary index of fineness, (b) ch = cotton hank (840 yd), (c) wr = worsted hank (560 yd),
(d) wo = woolen hank (1600 yd).
Materials uses Nec meaning English cotton count on some occasions, but ASTM
Standard D2260 uses cc instead, and D1907 uses N (these are rarely used in mills
throughout the world). The Textile Institute uses T for the direct system and N for the
indirect system, and a variety of subscripts are used to distinguish between a number
of subcategories.
A1.1.4 Conversion
In normal practice, it is unnecessary to go through a calculation each time a conversion
is required; generally a conversion factor can be used (Table A1.2). In the case of
converting from one direct system to another, one merely multiplies the known linear
density by the conversion factor. A similar procedure is used when converting from
one indirect system to another. When converting from indirect to direct, or vice versa,
then the factor must be divided by the known quantity. Referring to the use of Table
A1.2, an example is to convert from cotton count to tex. In this case, 590.5 must be
divided by the cotton count. These are known as cross-conversions. Another example:
to convert from cotton count to worsted count, multiply the given cotton count by 1.5.
A1.1.5 Plied yarns and examples of calculation
It is possible to make up a yarn by twisting together two or more finer yarns. The
process is called plying and the yarns are called plied yarns. To show that a yarn is
plied, it is normal to write both the yarn count and the number of plies separated by
a slash. In some systems, the quoted yarn number is that of each component. A 100
den/3 yarn would mean that 3 plies of 100 denier yarns were twisted together to give
320
Appendix 1
Table A1.2
Conversion factors
Direct
Indirect
To
tex
denier
cotton
count
worsted
count
woolen
count
metric
count
From
tex
denier
cotton count
worsted count
woolen count
metric count
–
0.111
590.5
885.8
310.2
1000
9
–
5315
7972
2791
9000
590.5
5315
–
0.667
1.905
0.591
885.8
7972
1.5
–
2.857
0.886
8.06
72.54
0.525
0.350
–
0.31
1000
9000
1.693
1.129
3.224
–
a yarn whose equivalent linear density would be approximately 300 denier. The
twisting causes a slight effect but this can be ignored for now. With a direct system,
one adds together the individual linear densities to arrive at the total.
The problem is a little more complicated with an indirect system. Ignoring twist
effects, imagine a number of cotton yarns lying side by side, each being of the same
length L yards. If the masses of the strands are w′, w″, w″′ etc., then the total mass
M = w′ + w″ + w″′ + etc.
But
w′ = w″ = w″′, etc. = L/840
M = L l + l + l + etc.
840 N ′
N ′′
N ′′′
[A1.3]
where N ′, N″, N″′, etc., are the counts of the individual component yarns. If the
equivalent count of the ply is NT, then:
NT =
L
840 M
l = l + l + l + etc.
NT
N′
N ′′
N ′′′
[A1.4]
[A1.5]
In words, the reciprocal of the count of the ply is the sum of the reciprocals of the
component strands.
As noted in Chapter 3, the equivalent count is used by some in commerce, with no
indication that the number quoted refers to the equivalent count. Thus, one finds yarn
counts being written as, say, 10/2, meaning 10equ/2, whereas the practice elsewhere is
to use the form discussed earlier (i.e. 20/2) for the same yarn.
Dealing with short-staple yarns, the practice is to quote the ply as, say, 20/2,
whereas with long-staple yarns the numbers are reversed (i.e. 2/20).
A1.1.6 Simple draft calculations
The flow through a draft zone obeys the law of conservation of mass flow. In other
words,
Mass flow in = Mass flow out + losses of mass
[A1.6]
Calculations I: Elementary theory
321
In mass flow, the element of time is introduced and velocity is substituted for length.
Thus Equation [A1.6] may be re-quoted in the form:
Vini = Vono + losses
[A1.7]
where V = velocity of the fibers and n = linear density.
Ignoring losses,
ni/no ≈ Vo/Vi
[A1.8]
Often it is assumed that the speeds of the fibers and the rolls in contact with them are
the same, but this is not always so, because of slippage. Also, it is sometimes assumed
that the linear density of the final output product is the same as that of the material
passing through the output rolls. This is not always true either, because there can be
shrinkage immediately following the emergence of the strand from the front drafting
elements. That is why the approximately equals sign appears in Equation [A1.8]. The
ratio of linear densities of the output product and the input is called the actual draft
ratio (often referred to as just draft). The ratio of the surface velocities of the media
inducing flow (such as rollers) is known as the mechanical draft ratio or mechanical
draft. When several stages of drafting are used, the overall draft across them is the
algebraic product of the stage draft ratios. Taking two stages, designated by the
subscripts 1 and 2, and realizing the output of stage 1 is the input of stage 2, then no1
≈ ni2 and Vo1 ≈ Vi2. Let n ≈ no1 or ni2 and V ≈ Vo1 or Vi2.
Thus, since:
ni1/n ≈ Vo1/Vi1 and n/no2 ≈ Vo2/Vi2,
(ni1/n) × (n/no2) ≈ ni1/no2 ≈ draft cross the two stages
If no1 and ni2 are used, the drafts would give a close approximation to the actual
drafts. As mentioned, these are defined by the ratio of linear densities of the strand
at the input and output of each stage. However, in practice, spot measurements of
linear density of material moving through the system are rarely completely representative
and accurate. For the purposes of simple mill floor calculation it is normal to ignore
the losses, slippages and contractions, which are quite small. In such cases, normal
equals signs may be used as in Equation [A1.9]. The drafts so calculated are mechanical
draft ratios.
Thus, if ni1/n = ∆1 = draft in stage 1, and n/no2 = ∆2 = draft in stage 2
ni1/no2 = total draft = ∆ = ∆1 × ∆2
[A1.9]
This can be extended to total draft = ∆1 × ∆2 × ∆3 × ∆4 etc.
A1.2
Yarn diameter
A1.2.1 Diameter and cross-sectional shape of a yarn in service
It might be thought that the obvious way to describe a yarn would be by its diameter,
but there are difficulties with this approach. A textile yarn, by its very nature, has to
be soft and can squash; therefore, although it is approximately round in cross-section
when it is in the free state, it rarely remains round in fabric form. Different fibers are
used in all sorts of combinations and it is a complex matter to calculate fabric weights
322
Appendix 1
because of the physical differences in the fibers and fabric structures. Nevertheless,
it is helpful at times to have an idea of yarn diameter. For example, the diameter helps
determine how closely the yarns can be packed to make a fabric, or how well a given
yarn will cover in a given fabric.
A1.2.2 Theoretical diameter of a yarn in the free state
Let the linear density of a yarn, ny, be equal to the product of the number of fibers,
m, and the average linear density of the fibers, nf. Cover is the percentage area
covered by one or more yarns as they lie in the fabric.
n y = m × nf
[A1.10]
Assume that the fibers are evenly spread throughout the cross-section at a rate of b
fibers per square inch. A round yarn, of diameter d inches, has a cross-sectional area
of Ay, but Ay = πd 2/4 sq inch, and the yarn contains Ay b fibers.
From Equation [A1.10]
m = n y /nf, but also m = Ayb
substituting for Ay,
m = πd2b/4
substituting for m,
n y /n f = πd2b/4
whence
d2 =
4 ny
π bn f
and d =
4 ny
π bn f
[A1.11]
In a normal yarn, nf is relatively constant, but b is determined by the twist and
structure of the yarn. With a given yarn with a fixed twist and structure, the only
highly significant variable in the right-hand side of Equation [A1.11] is ny. Consequently,
we may write the value of the equivalent yarn:
d≈K
n
[A1.12]
The approximation sign is meant to take care of the uncertainties due to changes in
yarn shape; K is often treated as a constant factor, but care has to be taken in
exercising this option.
A1.3
Twist multiple calculations (staple spinning)
A1.3.1 Derivation of twist multiple as a function of yarn count
The helix angle at which the fibers lie is important in determining the properties of
the yarn. Since there is a profusion of helix angles involved in a yarn structure, it is
normal to define twist by the helix angle of the fibers in the outermost layer. This
angle is very roughly 45° for a normal yarn.
Calculations I: Elementary theory
323
A simple experiment will demonstrate the relationship between twist angle just
discussed (i.e. helix angle) and diameter. Draw a line diagonally on a flexible transparent
sheet and roll the sheet tightly. The line now appears as a helix consisting of numerous
repeats along its length. Allow the roll to increase in diameter and it will be seen that
the number of repeats decreases. Let the roll grow to such a diameter that there is
only the single complete helix as shown in Fig. A1.1. If the diameter of the roll is now
D, the length of a repeat (or wavelength) is λ, and the helix angle is β, then tan β =
πD/λ. But tan β is proportional to twist multiple.
Therefore
TM = KπD/λ
[A1.13]
where K is a constant.
A twisting or spinning machine has to be set to give a certain number of turns per
inch (say τ) but, if the yarn does not change in length, this is really the number of
helical repeats in one inch of yarn. If λ is measured in inches, 1/λ = τ (i.e. λ is the
number of inches per turn). Yarn diameter is roughly proportional to the square root
of the linear density (or to the inverse square root of the yarn count, according to the
system used). Take an indirect system, where K1 is a constant and N is the count, D
≈ K1/√N. Substituting in Equation [A1.13] and transposing:
Twist density = TM √N × constant, and the twist level is measured in turns/unit
length.
The constant is normally taken as unity and the twist is given by:
Twist density = TM √N tpi
[A1.14]
In this case, the twist is measured in turns per inch and this is usually contracted to
πD
β
λ
D
λ
Fig. A1.l
Geometric development of the strand surface
324
Appendix 1
tpi. However, it must be emphasized that ‘tpi’ is not the name of the variable but
merely describes the units of measurement. The variable is called ‘twist density’ and
we use the symbol τ to denote it. The result of applying the formula in this case yields
an answer in turns/inch. A subscript is added to N in Equation [A1.14] according to
whether the measurement system is English cotton count, worsted count, metric
count, or another indirect system.
A1.3.2 Twist multiple as a function of linear density
In the direct count system, similar logic produces the relationship:
τ = constant × TMdirect/N
[A1.15]
or
twist density = α/n
The factor α is frequently used in Europe in place of TM.
An ASTM standard recommends using a twist density measured in turns/cm and
assumes a constant of unity.
A1.4
Productivity of pre-spinning preparation machinery
A1.4.1 Opening line productivity
A normal opening line is capable of producing a fiber stream of the order of
1000 lb/hr and there are usually at least two opening lines in operation in a mill. Thus,
the minimum fiber stream found is of the order of 1 ton/hr. Let the minimum opening
line productivity be Po. (It might be noted that 1 long ton (UK) = 2240 lb and this is
approximately the same as 1000 kg (i.e. 2206 lb). The short ton (US) is 2000 lb.) The
efficiency of the process, η, is sometimes measured per unit (pu) rather than as a
percentage. Processing is carried out in a series of sequential stages. Machines following
the opening line have much lower productivities and there have to be multiple parallel
paths within a single stage of the later processes. The production should be approximately
matched at all stages and it is necessary to calculate the number of each sort of
machine required at each stage. The following calculations will be based on the
assumption that all machines at a given stage have the same productivities and that
the throughput is constant from stage to stage. This is not always true, but the calculation
will lay out the basic rules. It is not a great step to modify the procedures to accommodate
variations in the machinery mix.
A1.4.2 Sliver productivity
Cards, drawframes, combers, all produce a stream of non-twisted sliver and their
production rate is the mathematical product of the linear density and the delivery
speed. The productivity, P, is usually measured in lb/hr or kg/hr, the velocity, V, in
yd/min, or m/min and the yarn count, n, in grains/yd or g/m; the constant K has to be
adjusted accordingly and the efficiency, η, is expressed in per unit terms. Let the
machine productivity be P.
P = KVnη
[A1.16]
Calculations I: Elementary theory
325
A1.4.3 Card productivity
The productivity of a card is usually measured by direct weighing and the result
is expressed in lb/hr (or kg/hr). A typical figure, after losses, in the early 2000s
may well be in excess of 300 lb/hr (say 150 kg/hr). Let this productivity be Pc is
100 lb/hr. If the cards run continually, there have to be at least Po /Pc cards operating
in parallel, to match the prescribed Po output of the opening line. Prudence might
suggest a spare to allow for maintenance but that would be expensive. Consequently
for the figures suggested, we would require either [1 + (Po/Pc) = 1 + (2000/300)] or
[2000/300] depending on the risk that could be tolerated. These results may be
rounded up to 8 or 7 respectively (but we could just as well produce 2100 lb/hr with
7 or 6 cards) and speed up our opening line a bit. If production is stopped between
can changes, efficiency drops and the number of cards required increases in inverse
proportion to the per unit efficiency. If it is desired to reduce the throughput of the
cards to improve on nep performance, or to prolong the time between maintenance
stoppages, then a further increase in the number of cards is required.
A1.4.4 Drawframe productivity
Drawframes can deliver sliver at up to 1000 yd/min, even if normal production rates
are somewhat lower. The linear density of the sliver usually varies between 60 and
100 grains/yd. To establish an order of magnitude of drawframe productivity, let us
assume the sliver weight is 70 grains/yd (which is equivalent to 100 yd/lb), the
delivery speed is 800 yd/min, and an automatic can changer is in use (which implies
a per unit efficiency close to 1.0): also assume that the efficiency is 100%. Let the
machine productivity be P.
800 yd 70 grain
lb
×
×
× 60 min
yd
min
7000 grain
hr
= 480 lb/hr
P=
[A1.17]
Notice how the unit ‘yd’ cancels top and bottom, as do the units ‘grains’ and ‘minutes’,
and we are left with lb/hr. The result shows that 4.17 breaker drawframes are needed
to match the stated opening line output, and a similar number are required for subsequent
stages of drawing. The resulting figures have to be rounded up; therefore the number
needed would be at least 5 per passage. At least 10 drawframes would be required to
match the throughput of the opening system if two passages of drawing were used.
A1.4.5 Roving productivity
With roving, the product is twisted and the production is limited by the twister speed.
Flyer speeds can range up to 1000 r/min, and the twist levels are typically 0.9 tpi;
hence the linear speed is 1000/0.9 inches/min. (The value 0.9 is used rather than a
round number to facilitate tracking the calculation.) The linear density is often measured
in hank roving, where 1 cotton hank contains 840 yd of roving. A 1.2 cotton hank
roving contains 840 × 1.2 = 1008 yd/lb. The productivity equation2 may be modified
in form to give Pr = K′Vη/Ne.
2 K′ is a constant, Ne is the count, τ is the twist density, V is the linear velocity of the roving, and
U is the rotational velocity of the flyer. Alternatively, Equation [A1.24] may be used by converting
τ into TM.
326
Appendix 1
Also, V = U/τ and substituting for V we get:
Pr =
K ′Uη
τNe
[A1.18]
Thus, if U = 1000 r/min, τ = 0.9 tpi and Ne = 1.2 cotton hanks/lb (these are common
values for roving), a single spindle working without stop would produce:
yd
lb
Pr = 1000 × inches ×
× hank ×
min
0.9
1.2 hank 840 yd
36 inches
× 60 min = 1.837 lb
hr
hr
[A1.19]
Again, the units cancel to leave the final units as lb/hr.
If the efficiency were 0.9, then the net production would be 1.653 lb/hr and 1210
roving spindles would be needed per opening system capacity of 2000 lb/hr. The
result has, of course, to be rounded up to find the number of machines needed. Each
machine has a number of spindles specified by the machinery maker.
A1.4.6 Roving wind-on speed
The winding-on speed is Uw = ± (Uf – Ub ) r/min, according to whether the flyer or
bobbin leads. Some designs have the bobbin rotating faster than the flyer, in which
case Uw = (Ub – Uf ). The general case may use the absolute value, which takes no
account of sign, and it is written as |(Uf – Ub)| in the equation. When U is the speed
in r/min and r is the radius of wind in inches (the subscripts refer to the flyer and
bobbin), the linear velocity is:
Vw = 2πr|(Uf – Ub)| inch/min
[A1.20]
The roving supply velocity is Vs.
Vs = Uf /τ inch/min
[A1.21]
but Vs = Vw, therefore
2πr|(Uf – Ub)| = U f /τ
[A1.22]
But 2πrτ is the number of turns of twist put in a single coil of roving of radius r; let
this number of turns of twist be τc. Substituting τc in Equation [A1.22] and rearranging,
the ratio between the bobbin and flyer speeds is:
Ub
= 1± 1
τc
Uf
where Vw
Uw
Uf
Ub
r
τ
||
=
=
=
=
=
=
=
[A1.23]
winding-on speed in inches/min
winding-on speed in r/min
flyer speed in r/min
bobbin speed in r/min
pitch radius of outer layer of roving in inches
twist density in twist/inch
means that + or – sign of the resultant within the ‘bars’ should be
ignored.
Calculations I: Elementary theory
A1.5
327
Ring frame performance
A1.5.1 Ring frame productivity
The productivity equation for the production of a twisted strand shown in Equation
[A1.18] can be modified further. In ring spinning, the twist multiple is the factor
most likely to be kept fairly constant, irrespective of yarn count. If we insert the
appropriate value of K′ in Equation [A1.18] we get:
P=
Uη
Uη
or P =
1.5
504 TM [ N e ]
504 TM N e N e
[A1.24]
With a typical count of 24 Ne and a TM of 3.5 being spun on a spindle rotating at
15 000 r/min, with an efficiency of 0.97, we get a productivity of 0.0701 lb/hr. We
would require at least 28 531 ring spindles to match our hypothetical 2000 lb/hr
opening line. If a machine contained 800 spindles, then we would require at least
35.66 machines. Rounding this up, a practical number would be 40 machines. The
requirement for spindles also changes if the average count of the mill changes. An
operator running with a bare minimum number of spindles would encounter a shortage
of spinning capacity when the average count goes finer. If the average count goes
coarser, there will be some surplus spindles. In practice, the number of spindles
cannot easily be changed and the speeds are manipulated to balance the system as far
as possible. It should be mentioned that the average count and twist are usually
determined by the marketplace and the spinner has limited choice in the matter.
A1.5.2 Elements of mill balance
The flow through the various production phases has to be balanced if the machines
are to be fully utilized. The number of each sort of machine can be calculated from
the quotient of total mass throughput divided by the productivity of the particular
machine. Adjustments have to be made for fiber losses from each stage and the
proportion of those losses recycled. An example of such a calculation is given in Q34
in Appendix 2. In addition, where dissimilar machines work in parallel, it is necessary
to calculate the production of each of the parallel production streams and add the
results to make a balance with the overall product flow.
A1.5.3 Bobbin flow
A spinning bobbin might only contain 0.1 or 0.2 lb of yarn; thus, a large number of
bobbins has to be handled. In our hypothetical case, the mill operator would have to
handle at least 10 000 bobbins per hour. This highlights why attention has been given
to automating the transfer between the ring frame and the winder.
A1.6
Winding performance
A1.6.1 An example of the reduction in winder productivity
due to the need to splice
A winding machine might run up to 1000 yd/min and, if it were not for the interventions
needed to perform its clearing function, the productivity when winding a 6s yarn
328
Appendix 1
could be in the order of 12 lb/hr (0.2 lb/min). If, for example, a 5 lb cheese of yarn
has 30 splices in it, and each splice needs 10 seconds to locate the fault and perform
the splice, then 5 minutes of winder time is spent on splicing during the build of the
package. If it had not been for the splices, our hypothetical winder would have taken
25 minute to wind the yarn. For simplicity, assume that the time to doff the full
packages and replace them by empty ones is negligible and assume there is no loss
of fiber through the various processes. The total winder head time consumed would
have been 30 minutes and the average speed would have fallen to 10 lb/hr.
A1.6.2 An example of the effect on winder productivity caused by
poor yarn quality
If the yarn fault rate mentioned in Section A1.6.1 doubled and 10 minutes of winder
time was on splicing, the average speed would drop to 5 lb/35 min = 0.143 lb/min,
which is equivalent to 8.57 lb/hr. Thus, the number of winders required depends not
only on the count of the yarn, but also on its quality. In the best of the two cases just
discussed, we would have needed at least 14 winding spindles to deal with the production
rates of our hypothetical mill. In the case of the higher fault rate, we would have
needed at least 21 spindles.
A1.7
Rotor spinning machine performance
A1.7.1 General statement about productivity in rotor spinning
Rotor spinning needs no roving or winding facilities and the preparation of sliver is
much the same as that already discussed. The biggest difference is in the speeds that
can be attained. Rotor speeds of up to 130 000 r/min are possible, and preparation has
to be good enough to prevent rotor fouling and provide reasonable long-term evenness.
Equation [A1.24] applies.
A1.7.2 An example of rotor productivity
Appropriate data in this case might be: rotor speed = 100 000 r/min, spinning efficiency
= 0.97, TM = 4.0, and yarn count = 24s. If we apply Equation [A1.24], the productivity
is about 0.4 lb/hr and at least 5000 rotors would be needed for our hypothetical mill.
Appendix 2
Calculations II: Worked examples
A2.1
Yarn numbering
Q1. How many yards of yarn are there in 1 lb of 24s cotton yarn?
Ans. In 1 lb of cotton yarn there are 840 × 24 = 20 160 yd. In other words, there are
840 Ne yd/lb in the cotton system of units.
Q2. If there are 20160 yd in 1 lb of yarn, what is the worsted count?
Ans. L/W = 560 × Nw
where
L = length in yd and W = weight in lb
Nw = L/(560 W) = 20 160/560
= 36 worsted hanks/lb or Nw = 36s
From the foregoing, it will be realized that Nw is 50% larger than Ne for the same
yarn.
Q3. If Ne = 24s, what is the linear density in g/m?
Ans. This is a conversion from an indirect system to a direct system; therefore, one
is inversely proportional to the other. In other words n′ = constant × (1/Ne), consequently
Ne should appear on the bottom line of the equation.
There are 1.093 yards in 1 meter and 454 grams1 in a pound.
A 24s yarn has a length of 24 × 840 yd/lb and, designating gram and meter by g
and m, respectively:
n ′′ =
1lb
454 g 1.093 yd
×
×
(24 × 840) yd
lb
m
n″ = 0.0246 g/m
Q4. What is the linear density of the yarn in Q3, expressed in tex?
1 The French spelling, gramme, is often used, to avoid confusion with grain, especially when hand
written.
330
Appendix 2
Ans. The unit tex is the same as g/km.
1000 m
n ′′ g
×
km
m
= 1000 × 0.0246 g/km = 24.6 tex
n=
Q5. If a 120 yd skein weighs 40 grains, what is the cotton count?
Ans. We have to change the units. Remembering that there are 7000 grains in 1 lb,
and 840 yd in 1 cotton hank,
Ne =
7000 grain 1 cotton hank
120 yd
×
×
40 grain
lb
840 yd
Canceling out the units as well as the numbers, we get:
Ne = 25 cotton hanks/lb
A2.2
Drafting
Q6. A roving of 1 hank roving (Ne = 1) is converted to a 24s yarn. If twist contraction
is ignored, what is the actual draft ratio?
Ans. Actual draft ratio = (output value of N)/(input value of N) in compatible units.
Nei = 1.0 cotton hanks/lb and
Neo = 24 cotton hanks/lb
Actual draft ratio = 24.
Q7. The linear velocity of a yarn leaving a drafting system is 100 ft/min, and the
entering material has a velocity of 2 ft/min. What is the mechanical draft?
Ans. Mechanical draft = Vo /Vi = 100/2 = 50.
Q8. A roller drafting system consists of two pairs of drafting rollers; the front rollers
are 1 inch diameter and the back rolls are 1.25 inch diameter. The front rollers rotate
at 90 r/min and the back rollers at 3 r/min. The system is fed with 2 hank roving (Ne
= 2 cotton hanks/lb). What is the yarn count if twist contraction is ignored?
Ans.
Vo = πDoU = 90π inches/min
Vi = πDiU = 3 × 1.25π inches/min
Mechanical draft = VoVi = 90π/3.75π = 24
Output Neo = 2 × 24 = 48 cotton hank/lb.
Q9. If the yarn delivered in Q8 contracts by 3% before it is wound, what is the actual
draft ratio?
Ans. Without shrinkage (output Neo) = 48 × (input Nei).
After shrinkage, the yarn is fatter and the output Neo is less, thus the actual draft ratio
= 48 × (1.00 – 0.03) = 46.56 cotton hanks/lb.
Q10. A sliver-to-yarn drafting system is fed with 50 grain/yd sliver and delivers a
strand of Ne = 24 cotton hanks/lb. What is the actual draft?
Ans. The input is expressed in a direct system and the output in an indirect one. Thus,
the first step is to convert one value into the units of the other, because compatible
units must be used.
Calculations II: Worked examples
Input N e =
331
7000 grain 1 hank
1 yd
×
×
= 0.1666 cotton hanks/ lb
50 grain
1 lb
840 yd
The second step is to state in the input and output counts.
Output Neo = 24 cotton hanks/lb
The third step is to check the compatibility of the units and the fourth step is to
calculate the ratio as follows:
Actual draft = 24/0.1666 = 144
Q11. A drawn filament bundle is made up of filaments of 1.5 denier (i.e. 1.5 dpf).
The draw ratio used to orient the molecular structure was 5. What was the denier of
the original ‘spun’ filaments before drawing?
Ans. Output linear density = no = 1.5 denier.
Input linear density = no × draw ratio = 1.5 × 5 = 7.5 denier.
Q12. A 150 denier yarn is made of 1.5 dpf fibers. How many fibers are there in the
cross-section?
Ans. No of fibers in cross-section = nyarn/nfil
= 150 denier/1.5 denier
= 100 filaments/yarn.
Q13. A toothed drafting system takes in sliver at 53 grains/yd and converts it to a
stream of fibers that average 5 fibers in the cross-section. The fibers have a linear
density of 1.5 dtex. What is the draft ratio?
Ans. Linear density of input = ni
ni =
53 grain
1 lb
454 g 1.09 yd 1000 m
×
×
×
×
yd
7000 grain
m
lb
km
= 3747 g/km
Since 1 tex = 1 g/1000m, linear density of input = 3747 tex, and of output = 5 ×
1.5/10 = 0.75 tex. The units of input and output are compatible, hence
Draft ratio = 3747 tex = 4996
0.75 tex
Q 14. The foregoing is a very high draft ratio, typical of these devices. What is the
draft when the thin stream of fibers is condensed into a 30 tex yarn? What is the
overall draft?
Ans. The new input linear density for the second stage is 0.75 tex and the output is
30 tex. Therefore the draft ratio is 0.75/30 = 1/40. In other words, the condensation
stage gives a fractional draft. The overall draft is 3750 tex/30 tex = 125.
Q15. Four 40s yarns are plied. What is the equivalent count of the plied yarn if twist
effects are ignored?
Ans.
1 = 1 + 1 + 1 + 1 = 4
40
40
40
40 40
NT
from which it follows that NT = 10 hanks/lb.
332
Appendix 2
Q16. A 40s yarn is plied with a 20s and a 10s yarn to make a fancy yarn. What is the
equivalent count, if twist effects are ignored?
Ans.
1 = 1 + 1 + 1 = 1 + 2 + 4
40
20 10 40
40
NT
40
= 0.025 + 0.05 + 0.10 = 0.175 (hank/lb)–1
This is the reciprocal of NT, hence
NT = 1/0.175 = 5.71 hanks/lb
This calculation is typical of all indirect systems. To apply it to a particular one, make
sure to quote what sort of hank is involved. For example, if this had been wholly in
the cotton system, the answer would have been quoted as 5.71 cotton hanks/lb.
However, if it had been in the worsted system, the answer would have been 5.71
worsted hanks/lb. If the counting systems had been mixed, the answer could have
been expressed in one of the systems but the units used in the calculation would have
had to be consistent with the answer. Notice how the equivalent yarn count is smaller
than that of any of the component yarns.
A2.3
Belt transmission
Q 17. Consider a belt or yarn being driven in the direction shown in Fig. A2.1, by a
pair of rolls, one of which is 1.3 inch radius and it rotates at 110 r/min. The linear
velocity of the yarn, V, equals ωr. What is the velocity? (Hint: care has to be taken
with the units. If V is to be in ft/sec, then the rotational speed, ω, must be expressed
in radians/sec and r in feet.)
V = 1.25 ft/sec
110 rpm
r = 1.3 inch
Fig. A2.1
Strand delivery
Calculations II: Worked examples
333
Ans.
V = ωr
ω = 2 π × 110 rev × min rad/ sec
min
60 sec
ft
12 inch
r = 1.3 inch ×
ft
V = 2 π × 110 rev × min ×
× 1.3 inch
min
60 sec 12 inch
= 1.25 ft/sec
Q18. Determine the speed ratio of the pulleys shown in Fig. A2.2(a). The large pulley
has a radius of rL inches the small pulley of rs inches and they rotate at UL and
Us r/min, respectively.
Ans. The belt speed can be determined by considering either the small pulley or the
large one. The belt thus runs at:
V = KULrL ft/sec, where K = 2π/(60 × 12)
The small pulley radius is rs and it rotates at Us r/min, which gives:
V = KUsrs ft/sec
KULrL = KUsrs
Thus
ULrL = Usrs
or
ULDL = Us Ds
[A2.1]
where D = diameter and the subscripts have the same meaning as already explained.
It will be noticed that the constants cancel because we are dealing with ratios.
Q19. An electric motor runs at 1800 r/min and drives a shaft by a pulley and belt
US
UL
V
rs
(a)
US
rL
UL
V
rS
(b)
Fig. A2.2
rL
Belts and pulleys
334
Appendix 2
system. The pulley on the motor is 6 inches diameter and the pulley on the driven
shaft is 18 inches diameter. What is the speed of the driven shaft?
Ans.
Us = 1800 r/min, UL = ? r/min
Ds = 6 inches, DL = 18 inches
From Equation [A2.1], UL = Us Ds/DL
= 1800 × 6/18 = 600 r/min
This answer is not completely accurate – see Q20 (b) and (c).
Q20. (a) What would be the effect if the belt of Q19 is crossed? (b) What effects
would slippage have? (c) What effect does belt thickness have?
Ans. (a) If the belt were crossed as in Fig. A2.2(b), the direction of rotation of the
driven member would be reversed and a minus sign can be introduced to take this into
account. Thus the answer for the crossed belt case is minus 600 r/min.
Ans. (b) There is always a slight amount of belt slippage, which slightly reduces the
speed of the driven member.
Ans. (c) The thickness of the belt cannot be ignored. It is usual to add one belt
thickness to the actual pulley diameters in calculating the speeds. If a 1/8 inch thick
belt were used in the foregoing example, and slip is ignored, the approximate speed
would be:
U1 ≈ 1800 × (6.0 + 1/8)/(18 + 1/8)
≈ 1800 × 6.125/18.125
≈ 608 r/min
A2.4
Gearing
Q21. A motor runs at 720 r/min and drives a shaft by means of a sprocket and chain.
The motor sprocket has 20 teeth and the driven sprocket has 80 teeth. What is the
speed of the shaft?
Ans.
Let output speed = Uo
Output speed = 720 × 20 = 180 rpm
80
Q22. A compound gear system consists of a 20 tooth driver that meshes with an 80
tooth gear and the latter is locked concentrically with a 25 tooth gear that meshes
with the output gear as shown in Fig. A2.3. The gear ratio is 16:1. How many teeth
are there in the output gear?
Ans. Let the output gear have m teeth.
Gearing ratio = (–80/20) × (–m/25) = 16
whence m = 16 × 20 × 25/80 = 100 teeth.
A2.5
Machine speeds
Q23. To be able to get a reasonable output per card and yet only have a thin web of
fibers on the main cylinder, it is necessary to have a high surface speed. Suppose
Calculations II: Worked examples
335
80 teeth
20 teeth
Output
gear
25 teeth
Mesh
Mesh
Fig. A2.3
Compound gears
there are 200 fibers/sq inch on the surface of a 40 inch wide card. There are 200 × 40
= 8000 fibers/inch of circumference on the card. If a single cotton fiber weighs 1.3
× 10–8 lb, there are roughly 8000 × 1.3 ×10–8 = 10.4 × 10–5 lb/inch of circumference.
Assuming an output of 100 lb/hr, what is the surface speed?
Ans.
1 inch
vo = 100 lb ×
× 1 ft × 1 hr
–5
hr
10.4 × 10 lb 12 inch 60 min
= 1335 ft/min
Q24. What is the rotational speed of the cylinder in Q23 if the diameter is 40 inches?
Ans. U = V/πD r/min.
The diameter concerned must be expressed in feet to be compatible with the
velocity in ft/min. The diameter is 3.333 ft, V = 1335 ft/min, and U = 127 r/min.
Q25. If the output is to be 65 grains/yd sliver, what is the sliver delivery speed in
Q23?
Ans.
1 yd
3 ft
7000 grain
V d = 100 lb ×
×
×
× 1 hr
hr
65 grain
lb
yd
60 min
= 538 ft/min (or 179 yd/min)
A2.6
Twist calculations
Q26. What is the twist density, in tpi, of a 4 TM, 25/1 cotton yarn?
Ans. From Equation [A1.14], twist density = TM √Ne = 4 √25 = 20 tpi.
Q27. A 20 tex yarn has a TMdirect of 36 (α = 36); what is the twist level?
Ans. Twist level = 36/√20 = 8.05 turns/cm.
Q28. A yarn has 20 tpi and a count of 36s in the cotton system. What is the twist
multiple?
Ans. Twist multiple = TM = 20 / √36 = 3.33. No units need be quoted in this case.
336
Appendix 2
Q29. Plot a graph of twist level versus count, for a TM of 3.0.
Ans. Set out a table of co-ordinates.
Table A2.l
Co-ordinates of graph
4
2.0
6
Ne
Ne
τ tpi
9
3.0
9
16
4.0
12
25
5.0
15
36
6.0
18
The data are plotted in Fig. A2.4.
20
Twist/ inch
TM = 3
10
0
0
10
20
30
Yarn count (Ne)
Fig. A2.4
A2.7
40
Twist characteristics
Production
Q30. The front roll of a drafting system advances a strand into a twister that rotates
at 10 000 r/min. The roll diameter is 1.2 inch. Calculate the front roll speed when a
yarn of Ne = 25 hanks/lb and TM = 3.5 is being made. What is the speed ratio between
the spindle and the front roll?
Ans.
Twist density = τ
τ = TM N e
= 3.5 25 = 17.5 tpi
Ut = twister speed in rev/min
V = linear speed of yarn in inches/min
= U t /τ
V=
10 000 rev
× inch
min
17.5 turn
= 571 inch/min
Calculations II: Worked examples
337
But V = Ufr πD
where Ufr = rotational speed of front roller and D = diameter of front roller.
Substituting for V and D and rearranging:
rev
U fr = 571 inch ×
= 151 rev/ min (i.e. r/ min)
min
1.2 π inch
Velocity ratio = 10 000/151 = 66.
In other words, the spindle has to rotate 66 times as fast as the front roll of the
drafting system.
Q31. A roving frame running at 1000 r/min and producing a 1.1 hank roving
(Ne = 1.1) at 0.9 TM will produce P lb/spindle hour. What is the value of P?
Ans. Assuming the pu (per unit) efficiency is 1.0
P=
P=
Uη
504 × 0.9 × 1.1 ×
1.1
1000
504 × 0.9 × 1.1 ×
1.1
= 1.91 lb/spindle hr
[A2.2]
Q32. A traveler slides at 120 ft/sec on a 1.75 inch diameter ring. The twist density of
the yarn being spun (τ) is 20 tpi and it is wound on to a 1.25 inch diameter bobbin.
(Figure 7.3 shows a ring and traveler.) What is the percentage difference between the
traveler and package speeds? What does this difference represent?
Ans.
Let ωt = rotational speed of the traveler, and since ωt = V/R:
2
ω t = 120 ft ×
× 12 inch = 1645.7 rad/ sec
sec
1.75 inch
ft
Let Ut = rotational speed of traveler in traditional units
U t = 1645.7 rad × 60 sec = 15 718 r /min (i.e. rev/min)
2π sec
min
V = the linear speed of the yarn = U/τ
Since τ = 20 tpi, V = 15 718/20 = 785.9 inches/min
d = 1.25 inches and the wind-on speed = V/πd
Uwind = 785.9/(1.25π) = 200 r/min
This is 1.27% of the traveler rotational speed.
Ring spindle speed = Ut + Uwind = 15 718 + 200 = 15918 r/min
Note: As the bobbin diameter builds from (say) 1 inch to 1.6 inch, with a bobbin
speed of 15 918 r/min, the wind-on speed varies from
785.9/π = 250 r/min to 785.9/(1.6π) = 156 r/min
and the traveler speed varies from
15 918 – 250 = 15 668 r/min to 15 918 – 156 = 15762 r/min,
338
Appendix 2
a difference of about 0.6%. As the bobbin diameter changes, a small variation in twist
occurs but the effect of this is neglected.
Q33. A ring frame produces a yarn of average count of 25/1. The twist multiple is 3.5
and the spindle speed is 20 000 r/min with a spinning efficiency of 0.95. (a) What is
the output for the given ring frame? (b) If the count were reduced to 36/1, what would
be the output?
Ans. (a) Equation [A1.24] contains the group Ne√Ne and it is easier to calculate this
first.
Ne√Ne = 25 × √25 = 125
Substituting this in Equation [A1.24] we get:
20 000 0.95
×
× 1 = 0.0862 lb/ sp hr
504
3.5
125
(b) Calculating Ne√Ne as a preliminary step,
Ne√Ne = 36 × √36 = 216 and inserting this in Equation [A1.24] we get:
P=
20 000 0.95
×
× 1 = 0.0499 lb/ sp hr
504
3.5
216
At least 12 spindles are needed in one case, and 20 in the other, to produce 1 lb/hr.
P=
Q34. A mill has an output of 2500 lb/hr of yarn of 16/1 (Ne) at 3.8 TM spun on ring
frames running at 15 000 r/min at an efficiency of 0.92 and a waste level of 1.8%. The
ring frames are supplied with 1.1 hank roving (Ne), made on roving frames running
at 1200 r/min and with a TM of 0.996. The efficiency of the roving frames is 93% and
the fiber loss is 0.2%. (a) How many ring frame spindles, and (b) how many roving
spindles are required?
The mill has two passages of drawing and the drawframes run at 600 yd/min when
producing 90 grain/yd sliver. (c) How many drawframe heads are needed if the
operational efficiency is 95%, the sliver wastage is 1%, and each drawframe has two
heads?
It is intended to install cards, each with a productivity of 100 lb/hr. The waste fiber
from carding and opening is 2% and the operational efficiency is 96%. (d) How many
cards would be needed, (e) what input fiber flow would be required, and (f) what flow
of new fiber would be needed if 50% of the waste from spinning, roving, and drawing
is recycled?
Ans. (a) Starting this question with the ring frames, the yarn flow required from them
= 2500 lb/hr.
Calculating the value of Ne√Ne = 16 × √16 = 64, the productivity of one ring spindle
15 000 × 0.92
= 0.113 lb/ sp hr
504 × 3.8 × 64
The number of ring spindles needed = 2500/0.113 = 22 124.
If there were 800 spindles per machine, 27.66 machines would be needed; rounding
this up gives us 28 machines. (This number would have to be increased to allow for
maintenance shutdowns and repairs.)
(b) Allowing for 0.008 pu fiber loss in spinning, the roving flow needed is:
Prf =
2500 + (2500 × 0.008) = 2500 × 1.008 = 2540.16 lb/hr
Calculations II: Worked examples
339
TMroving = 0.996 and Ne √Ne = 1.1 × √1.1 = 1.154
1200 × 0.93
= 1.927 lb/ hr
504 × 0.996 × 1.154
Number of roving spindles needed = 2540/1.927 = 1318, say 1400.
(c) Drawframe production/head for one passage, Pdf = Vsliver × nsliver
Proving =
Pdf =
600 yd 90 grain
lb
×
×
× 60 min × 0.95 × 0.99
min
yd
7000 grain
hr
= 435.3 lb/ hr
Allowing for 0.002 pu (per unit) fiber losses in roving, the throughput is: (the value
in Answer (b) × (1 + 0.002)) = 2540 × 1.002 = 2545 lb/hr.
Number of heads required = 2545/435.3 = 5.847 for one passage.
For two passages the number required = 5.847 × 2 = 11.69 and rounding up, this
would be taken as 12.
With 2 heads/drawframe, 6 machines are required.
(d) Allowing for 0.01 pu fiber losses in drawing, the card output is:
(the value in Answer (c) × (1 + 0.01)) = 2545 × 1.01 = 2570 lb/hr.
Taking the efficiency into account, the production rate/card is 100 × 0.96 =
96 lb/hr. The theoretical number of cards required would be = 26.77. However, one
cannot have a fraction of a card so the number required is rounded up to 27.
(e) After losing 2% of the fiber in carding and opening, the input rate is:
2570 × 1.02 = 2621 lb/hr
(f) The specified wastes are:
Spinning waste = 0.008 × 2500
= 20.00 lb/hr
Roving waste = 0.002 × 2540.16 = 5.08 lb/hr
Drawing waste = 0.01 × 2570
= 25.71 lb/hr
—————
Total specified waste
= 50.79 lb/hr
Recycled waste = 50.79/2 = 25.4 lb/hr, which offsets the losses and the total fiber
requirement drops by this amount. Thus, the net input fiber required in this case is
2622 – 25.4 ≈ 2597 lb/hr.
A2.8
Texturing
Q35. A texturing machine has a six-disk stack. The coefficient of friction, µ, is 0.2,
the run-on and run-off angles are both 30°, and the inclination on the periphery of the
disk is 0°. Assume that all disks are working disks and that they give no aid in moving
the yarn through the stack. Calculate the output tension and the torque produced by
disks 1 and 6.
Ans. The angle of wrap for each disk = (90 – θ) × 2 = 120° (equivalent to
2π/3 radians). Let the input tension to disk 1 = T1 and, using Amonton’s Law (i.e. To
= T1 eµθ), output tension from disk 1 = T1 e0.2×2π/3 = 1.52 T1.
Since the passage past five disks accumulates an angle of wrap of five times that
of the passage over a single one, and the angle appears as an exponent in the equation
we may write:
340
Appendix 2
Input tension to disk 6 = 1.525 T1 = 8.11T1
Output tension from disk 6 = 1.526 T1 = 12.3T1
The torque generated by a disk = (T1 + T2) µ √n K cos θ
where n is the linear density of the yarn, µ is the coefficient of friction, and K is a
factor.
Torque generated by disk 1 is ((1 + 1.52) × 0.2 × 0.866) KT1 √n
= 0.436 KT1√n
Torque generated by disk 6 is ((8.11 + 12.3) × 0.2 × 0.866) KT1√n
= 3.54 KT1√n
This torque for disk 6 is 8.1 times that generated by disk 1, but the maximum tension
is 12.3T1; this is also 8.1 times the tension output of disk 1.
Appendix 3
Advanced topics I: Air conditioning and
utilities
A3.1
Introduction
Water vapor is very important in yarn making in many ways. Frequently, steam is
used as a heating medium. It is used because there is a unique relationship between
pressures and temperatures of steam, which makes it relatively easy to control. There
is often an advantage in applying moist heat to set polymers, especially when hydrogen
bonding occurs within the polymer molecular structure. The air in which we live is
really a mixture of air and steam. In a textile mill, where high humidities are needed,
the proportion of steam is rather high and it is helpful to understand its characteristics.
A3.2
Units
Conventionally in the USA, units are expressed in pounds, feet, and seconds, whereas
the SI system uses grams, meters, and seconds. The relationships between derived
units may not be obvious and it may be of use to discuss them. In thermodynamics,
the interest is in mass, temperature, and energy as well as some other parameters.
Temperature is commonly measured on two scales (i.e. Fahrenheit and Celsius) with
two important points being fixed by the freezing and boiling points of water under
normal atmospheric pressure. On the Celsius scale, water freezes at 0°C and boils at
100°C. For many thermodynamics calculations it is necessary to work from absolute
zero rather than from an arbitrary one fixed by the characteristics of a single substance.
This absolute zero is –273°C on the Celsius scale, which is awkward; it is preferred
to set absolute zero at the datum but use the same intervals as the Celsius scale. This
is now called the Kelvin scale. Conversion is made by merely adding 273 degrees to
the temperature in Celsius; the result is written as °K. The absolute zero measured in
Fahrenheit is –460°F; when the intervals are the same as the Fahrenheit scale, the
result is called the Rankine scale.
The heat content of a material is the product of mass × specific heat × temperature,
where the heat content and temperature are reckoned from some arbitrary levels. This
342
Appendix 3
assumes that there has been no change in state, such as from solid to liquid or from
liquid to gas. The specific heat is a property of the material; it is proportional to the
amount of heat that has to flow to or from a unit mass of the material to change its
temperature by 1 degree. In the conventional system, the specific heat of water is 1.0
and this is used as a reference for other materials. Temperature is a measure of
thermal ‘pressure’; heat will flow more rapidly through a material as the temperature
increases. Care should be taken to discriminate between temperature and heat. A
thermometer measures temperature but it is necessary to know the specific heat and
mass of a substance before the quantity of heat can be determined. In the conventional
US system, the unit of heat is the Btu and this is defined as that heat required to raise
one pound of water through 1°F. Heat is a form of energy but there are other forms.
For instance, electrical energy is measured in joules. Mechanical energy is measured
in ft lb (that energy needed to raise 1 pound through 1 foot). These are mutually
convertible; for example, 778 ft lb is equivalent to 1 joule. The Sl system expresses
all forms of energy in joules. Power is the rate of using (or producing) energy. In
engineering the common term is horsepower, which is defined as 550 ft lb/sec. In
thermodynamics, the term is often expressed in Btu/hr, whereas in electrical engineering
the unit is the watt. A watt can be variously defined as the energy flow in joules/sec
or the product of voltage and amperage. One ampere is the flow of electrical charge
in coulombs/sec; thus, since 1 coulomb may be defined as 1 ampere second, the two
definitions come to the same thing. All forms of power are mutually convertible and
the SI system uses the watt as the universal unit of power.
A3.3
Water vapor and steam
A3.3.1 Change of state
When a material changes from solid to liquid, liquid to gas, solid to gas, or any of the
reverse processes, we refer to a change of state. A change of state is nearly always
associated with a taking in or a giving out of heat. This is an important property of
most materials; water and thermoplastic polymers are no exception. Many of the
ideas discussed relate to water, but they have their counterparts with other materials.
Naturally, the values of the data can be widely different.
A3.3.2 The properties of steam
Water absorbs energy as it is heated, with the result that the temperature rises until it
reaches its boiling point. Beyond this boiling point, any further addition of energy
causes water to be converted to steam but the temperature does not rise again until all
the water has been boiled off. The boiling point rises with pressure. Table A3.1 shows
how the latent heat changes. It can be seen that the temperature of wet steam can be
controlled by the pressure in the steam vessel. With wet steam, some water remains
in the steam and, as energy is added, the percentage of water drops. When all the
water is converted, the steam is said to be ‘dry’. In practice, it is much easier to
control pressure than temperature, and, for this reason, the use of steam as a heating
medium is quite popular. Within limits, there is automatic regulation of temperature
all the time that the steam is wet and at constant pressure. If an autoclave (or steam
kettle) is used to set or dye yarn or fabric, it is usually necessary to adjust the kettle
to some pressure higher than atmospheric. For example, if a temperature of 150°C is
Advanced topics I: Air conditioning and utilities
343
required, the autoclave should be run at 70 psi or 482 kilopascals, as shown by Table
A3.1. The term ‘steam’ describes water vapor without any other gas or vapor present.
Steam that is just dry (i.e. the last drop of water has just been converted to steam) is
called saturated steam. A dryness fraction, q, is used to define the actual specific
volume of wet steam in respect to the specific volume of saturated steam at the same
pressure and temperature. The term ‘specific’ means that the quantity relates to unit
mass; specific volume is the volume taken by a unit mass of the substance. The
specific enthalpy or energy content of wet steam, H, is given by the simple formula:
H = s + qL
[A3.1]
where: s = specific enthalpy of water (i.e. sensible heat), L = specific enthalpy
associated with a complete change of state from water to dry steam (i.e. latent heat),
and q is the dryness fraction.
A3.3.3 The properties of air/steam mixtures
If two or more gases or vapors are mixed in a confined space and there is no chemical
reaction, each component fills the whole volume and each exists at its own particular
partial pressure. The sum of the partial pressures equals the total or applied pressure.
This is known as Dalton’s law of partial pressures. With a perfect gas, the mathematical
group P Vol/T remains constant for the particular gas and the constant is denoted by
the symbol G, which relates to energy/unit mass, and stands for the universal gas
constant. P is the pressure and Vol is the specific volume of the gas. The temperature
must be expressed in absolute terms.
With an imperfect gas or vapor, such as steam, the characteristic equation mentioned
is inaccurate and one must determine the specific volumes from tables such as Table
A3.1. In the case of steam/air mixtures, the absolute volume and temperature are
common to both the steam and the air. The mass of each component is the quotient
(absolute volume/specific volume). The absolute volume is the whole volume occupied
Table A3.1
Properties of steam
Conventional units
t
SI units
°F
P
psiabs
H
Btu/lb
Vol
cu ft/lb
L
Btu/lb
°C
t
P
kPa
H
J/g
32
40
50
60
70
80
90
100
150
212
250
293
320
358
417
0.0885
0.1217
0.1781
0.2563
0.3631
0.5069
0.6982
0.9492
3.718
14.696
30
60
90
150
300
1076
1079
1084
1088
1092
1097
1101
1105
1126
1150
1164
1178
1185
1194
1203
3306
2444
1703
1207
868
633
468
350
97.1
26.8
13.8
7.17
4.9
3.02
1.54
1076
1071
1066
1060
1054
1049
1043
1037
1008
970
945
915
895
864
809
0
4.4
10
15.5
21.1
26.7
32.2
37.8
65.6
100
121
145
160
181
214
0.61
0.84
1.23
1.77
2.5
3.49
4.81
6.54
25.6
101
207
413
620
1030
2070
2502
2491
2479
2465
2451
2440
2426
2412
2344
2256
2198
2128
2081
2010
1882
Vol
m3/kg
206.3
152.5
106.3
75.3
54.2
39.5
29.2
21.8
6.06
1.67
0.861
0.447
0.306
0.188
0.096
Notes: t = temperature, P = absolute pressure, H = total enthalpy of dry steam, Vol = specific volume, L = latent heat
344
Appendix 3
by the mixture. The specific volume of steam must be found from tables, whereas for
air, the characteristic gas equation can be used, i.e.
Pa Vol = MG
[A3.2]
where Pa is the partial pressure of the air, and Vol and M are the volume and mass of
air involved respectively.
There are two cases in which steam/air mixtures assume importance. One relates
to moist air, and the other to air leaks in autoclaves and boilers. In the second case,
the effect of air leakage into a steam system is to reduce the partial pressure of the
steam, which then causes a reduction in temperature. A normal gage can only measure
the sum of all the partial pressures and it cannot detect the displacement effect of the
intruding air. An inward air leak can cause a drop in temperature (which might be
undetected), and this can cause difficulty in some dyeing and setting operations. The
problem is compounded because water contains dissolved gases that are released on
boiling; the released gases act in the same way as the air leak. Thus, boiler water
should be de-aerated and steam traps should be used to permit removal of air and gas
without loss of steam. The dissolved air can cause corrosion in boilers and equipment
and it is prudent to remove it for this reason also. Air leaks are more likely where
steam cools and the internal pressure drops below atmospheric.
A3.4
Humidity
A3.4.1 Humidity in the workspace
Normal air is really a mixture of air and superheated steam. Superheated steam has
been heated above its saturation temperature, but if the partial pressure is very low,
the saturation temperature is also low. When the temperature of the humid air is
reduced to its dew point, the steam starts to condense and droplets of water are
precipitated to form a fog. In a workspace, water is deposited on cold surfaces and,
since these are often of steel or iron, there can be a problem with rust. Thus, it is good
practice to keep the temperature of the workspace above the dew point, always. As the
temperature of the air is increased above the dew point, the specific volume of the air
increases and the air is drier. The normal measure of wetness is relative humidity
measured on a 0 to 100% scale. It is, in fact, the ratio of the amount of moisture that
the air actually holds to the maximum that it could hold at the same temperature. It
can also be defined in terms of the partial pressures. Steam tables such as Table A3.1
could be used to calculate such conditions, but it is more normal to use psychometric
charts such as shown in Fig. A3.l. Two styles are given, one in SI (diagram (a)) and
one in imperial units (diagram(b)). To use these charts, one needs to know the wet
and dry bulb temperatures. These are measured by a pair of thermometers; one
element is kept dry and the other is kept moist. The evaporating water from the wet
bulb keeps the local temperature down to the dew point. It is possible to see how the
moisture content increases as the air is heated at constant rh by tracking the line AE
in diagram (b). It is also possible to see how the state point A is defined by (i) the wet
and dry bulb temperatures or (ii) the dry bulb temperatures and rh.
A3.4.2 Air conditioning
A mill has to have a controlled climate if high quality yarns are to be made. In
drafting, fibers can stick to the rolls and there are several possible causes for this
Air densities in m3/kg (shown with dashed line)
A = 0.80
B = 0.85
0.03
C = 0.90
30°C
Wet bulb
100
%r
h
80%
rh
60%
rh
40%
rh
0.02
20
10°C
Wet bulb
Dew point (°C)
C
20°C
Wet bulb
0.01
345
30
20
%
rh
Mass ratio = mass of moisture/mass of air
Advanced topics I: Air conditioning and utilities
B
10
A
10
20
30
40
Dry bulb temperature (°C)
%r
h
80%
rh
60%
rh
(a)
80
70
ra
rh
pe
tem
B
60
20%
lb
bu
50
60
40
Dew point (°F)
es
(° F
)
40%
rh
100
E
tur
100
W
et
Moisture content (grain/lb)
150
40
A
D
C
40
60
80
100
Dry bulb temperature (°F)
(b)
Fig. A3.1
Psychometric chart
problem. One is stickiness, caused by too damp an atmosphere. Another is electrical
charging of the fibers, which causes them to be attracted to a surface. The charge
becomes a problem when the atmosphere is too dry. The best rh of the air depends on
the fibers and the roll coverings, but a typical value is 55% rh.
346
Appendix 3
It is not a simple matter to get the rh to the correct value everywhere in the plant.
For example, in a spinning room, it is not unusual to find zones that are too wet or too
dry. Certainly, one must not rely on the wall-mounted hygrometers since they merely
record at fixed locations. The use of a portable hygrometer will quickly reveal the bad
zones. Hot areas, such as near a motor, give low values, and inappropriate values are
often found near doors, especially when the outside conditions are far from ideal. To
give an idea of the magnitude of the problem some mill experience will be quoted. In
one mill where the ring frames were positioned very close to one another, the rh at the
ring rail varied from 30% to 35%, even though the wall-mounted hygrometers read
55%. The mill performed badly. Another, with widely spaced ring frames and a welladjusted air conditioning system only showed ±1% rh. Most mills fall between these
extremes. The setting of the air diffusers and the pneumafil suctions can greatly
affect the uniformity of the rh throughout the room. A hint of poor distribution is
sometimes given by accumulations of fibers on the ceiling and light fixtures. Old
buildings with exposed beams and glassed areas are particularly difficult because of
the heat transfer through the roof and the large volumes of air trapped there. Careful
attention to air conditioning and distribution can save many later operational difficulties.
The ducting in the air distribution system must be designed to give uniform distribution
throughout the room. Also, the flows from the supply have to be balanced with the
main return air systems, as well as with the suction systems removing waste fiber.
A3.5
Mill environment
A3.5.1 Energy balance in an enclosed workspace
Considerable amounts of energy from electric motors and other devices are dissipated
in the workspaces. Not only is there heat dissipated from the motors, but also from
the machines themselves. The machines take mechanical energy from the motors and
do work in overcoming the resistance to movement of the machine parts and this
translates the energy into the heat form. Thus the machine parts get warm. For
example, bearings and belts get hot. The movement of the parts disturbs the air and
dissipates further energy; for example, the air leaving the rotors in OE spinning gets
very hot. The temperature difference between the machine parts and the air also
causes heat transfers to occur. Thus the original input electrical energy is translated
into heat energy at every stage in the process. In cool winter climates, heat escapes
from the entrances and exits as well as by conduction and radiation through the shell
of the building. Balance is normally obtained by applying a heating system. In hot
summer climates, the heat flows are reversed and air conditioning has to be applied
to keep the temperature down.
Consider the energy balance in an enclosed space:
If Eelec = electrical energy input to the space,
Ecomb = energy input derived from combustion of fuels
Etherm = thermal energy passed through the shell of the space due to a difference
in temperature between inside and outside.
Etherm = Total mass of enclosure × (Toutside – Tinside)
Tinside = is normally controlled to be constant and Etherm may have a positive or
negative value.
Eac = thermal energy pumped out of or into the space by the air conditioning
Advanced topics I: Air conditioning and utilities
Emat
347
plant. Eac is negative in summer when energy is being pumped out by the
refrigeration plant and positive in winter.
= is the sum of the differences between input and output in mechanical
strain, thermal and other energies resident in the textile material, which
differences are normally insignificant.
Since there is usually no chemical or nuclear reaction involved, the energies described
must be conserved and if similar units are used for all forms of energy:
Eelec + Etherm + Ecomb + Eac ≈ 0
[A3.3]
Availability of thermal energy is determined by the temperature at which it exists.
Every transfer degrades its availability. Eventually, it is all dissipated as low grade
heat that cannot be recovered economically. It is not just a question of balancing
Etherm + Ecomb and Eac but for every horsepower used within the space, there is a
dissipation of 746 watts from the machines themselves in addition to losses from the
electrical system. The energy from the lighting system is also dissipated as heat.
A3.5.2 Energy removal from the workspace
With a fixed amount of moisture present in the atmosphere, temperature changes are
accompanied by changes in rh. Consequently it is necessary to control both temperature
and rh. The additional heat from all these sources has to be removed to maintain an
even condition.
In the past, air conditioning often has been given little priority, with small regard
for the heat loading from the equipment within the building. However, as the equipment
installed consumes ever more power in the quest for higher productivity, the importance
of the heat loading becomes more apparent. In tropical or semi-tropical countries,
each kW of power used must be pumped out again while the refrigerating air conditioning
system is in use. This becomes particularly apparent with high speed rotor spinning.
Schemes where the motors and hot parts of a machine are cooled separately from the
main workspace have appeared and these are to be encouraged. This source of heat
might be useful in a cold climate but in a hot one it adds to the air conditioning costs.
Eckert [1] pointed out that a deciding feature of the air conditioning system for rotor
spinning is the increased direct exhaust air capacity of the machine itself. The high
temperature difference between the exhaust air and the ambient makes it easier to
pump out the heat. He quotes values of 27°C difference. Since that article, speeds
have risen and, because the energy consumed rises approximately to the 3rd power of
speed, the temperature difference he quotes must be low by modern standards. The
attractiveness of direct exhaust system cooling rises accordingly. Eckert also points
out that the distributions of supply, exhaust, and return air are of utmost importance
(as they are for any spinning operation). If the heat from the motor, head-, and
tailstocks is directly exhausted, it is then only necessary to remove from the spinning
room little more than the heat loading from the lighting and transmissions. This
means that the amount removed from the room at normal ambient temperatures is
less than one-third of the total. The remainder is removed at temperatures up to 40°C
higher. A refrigeration process is sensitive to temperature differences across its cooling
coils and this means that it is easier and cheaper to remove the heat from the hotter
air. Also, heat removal by water washing of hot air reduces the temperature more
quickly than it does with air at atmospheric temperature. Combinations of air-wash
348
Appendix 3
and refrigeration can be designed for optimum efficiency and the proportions for hot
and normal temperature returns may well differ.
Against the gains in operating cost have to be set the costs arising from the extra
capital investment needed. Ducts to carry away the hot air (preferably underfloor)
have to be installed. Separated air conditioning systems are also desirable to deal
with the two classes of return air. Obviously it is an advantage to install such systems
when the plant is first built.
A3.5.3 Filtration
Not only does the air in a mill have to be maintained at an rh best suited to the
particular task, but it has to be clean. In many regions of the world, air quality in mills
is quite stringently monitored. There are regulations in many countries mandating
maximum levels of particulate matter in the air in the opening and carding areas.
Such filtration is especially important in cotton spinning because of the incidence of
byssinosis (an allergic lung disease) in some workers. Even if automation is used to
minimize the amount of human exposure, the regulations still apply. In other forms
of processing, noxious chemicals can be given off and the climate has to be controlled
there too. A further important reason for attention to cleanliness is the fire risk. Some
airborne fibers and dust are flammable; fire and explosion risks are severe. Enclosed,
spark-free motors have to be used and a number of fire hazards are outlawed. Special
fire-fighting arrangements must be provided and most local authorities have a Fire
Code, which requires compliance.
Lint and harmful dust have to be removed from the return air. Where the
concentrations of dust and fly are high, such as in the returns from the carding and
opening areas, cyclone separators are used to extract the heavier fraction of waste.
Most of the remaining dirt is removed by electrostatic precipitators, fabric filters or
the like. The air is usually washed as well.
One important concern is the maintenance of the atmosphere within the comfort
zone of the operatives (Fig. A3.2), especially if maximum performance is expected.
Moisture content of air
(grain/ lb)
A3.5.4 Fly
Another concern arises about the level of fibers in the air (fly) because it can and
does cause defects in the yarn produced. An accumulation of fly landing on the yarn
being made in a ring frame obviously creates a blemish. Not so obvious is the result
of fly on other products. For example, fly landing on roving during manufacture is
100
100% rh
70% rh
30% rh
Human
comfort
zone
50
40
60
80
100
Temperature, dry bulb (°F)
Fig. A3.2
Human comfort zone
Advanced topics I: Air conditioning and utilities
349
often thrown off again during spinning, only to land on the yarn. A good spinner
keeps an eye open for the sources of fly production and tries to eliminate them.
Fiber and dust can carry electrical charges, and so can the surfaces of machinery.
Friction between moving surfaces, and separation of those surfaces, also produce
electrical charges. Much of the dust and fiber is highly flammable and if the electrical
charges build up sufficiently to cause a spark, then a dangerous conflagration can
occur. The remedies are to keep rh at a proper level and to ground all machinery by
connecting it to earth with a conductive cable. In that way, electrical charge build-up
is minimized and any that does form is leaked to earth before it can cause damage [2].
There are also other reasons for controlling the electrical charge. If the rh is too
low, electrified fibers coagulate and interfere with processing. The optimum level
varies from fiber to fiber and from process to process. A common symptom of
incorrect humidification in spinning is when roll laps occur.
A3.5.5 Lighting
Most workspaces in a modern mill are lighted exclusively with artificial lighting. The
electrical load can be limited to about 16 watt/square meter and still provide the
necessary 550 lux level of illumination. To achieve this economy, efficient light
sources, good reflectors, and clean, well-maintained, light-colored ceilings must be
used. Windows not only allow passage of light, but they form an easy path for noise
and heat transmission. Thus, natural lighting is avoided because of the increase in
load for the air conditioning plant and the increase in noise radiation to the outside.
Maintenance costs of the windows are also avoided.
A3.5.6 Effects of chemical contamination
Fibers or subsequent products treated with noxious chemicals can form a hazard.
Chemical emissions from any product or machine in the mill must be strictly controlled
and the necessary venting must be supplied. Such emissions are usually subject to
regulation by the local authorities. Where singeing is used, not only must the products
of combustion be properly vented, but the work area must be sealed off from the main
work areas to minimize the fire risks. The particulate level in the air should also be
monitored because soot inevitably escapes into the atmosphere.
In filament texturing, where the filaments are raised to high temperatures, lighter
fractions of fiber finish may boil off. It is important that the gases emitted are not
toxic, harmful to the product, or harmful to the machine. Again, proper ventilation is
required.
Man-made fibers have fiber finish which sometimes becomes removed from the
fibers and accumulates on certain important surfaces on various machines. Also
some fibers can produce oligomers which deposit on the working surfaces. In places,
the deposit forms a so-called ‘snow’ that is a sure sign of this sort of trouble. Cleanliness
in this respect is a necessary condition for the production of high quality products.
References
1.
2.
Eckert, O. Up-to-date Engineering design and Planning in the Concept of Open-end Rotor
Spinning Plants, Textile Machinery: Investing for the Future, Textile Inst Ann Conf, 1982.
Anon, Static Electricity, Nat Fire Protection Assn, Boston, USA, 1947.
Appendix 4
Advanced topics II: Testing of textile materials
A4.1
Divisions in testing
A4.1.1 Introduction
There is a gray area concerning quality control and testing. In one sense it is very
clear that the technicians who carry out the testing should have some say in the
sampling and interpretation of results. However, it is not always clear where the
measurements and techniques stop, and where the use of the results as a control
medium starts. The reason for testing is to acquire the data on which sound decisions
can be made; the reasons for quality control are to ensure adequate quality of product
and minimum trouble in processing. In a mill, testing is not an end in itself.
Testing falls into two categories: laboratory (or offline) testing and online monitoring.
In the first mode, shown in Fig. A4.1(a), samples of the product are taken to the
laboratory for testing. Various sorts of test apparatus are used. The test laboratory
must be air conditioned because the moisture content of textile fibers varies with
ambient conditions and many tests are strongly influenced by moisture content. Normally,
a laboratory is maintained at 70 ± 2°F (21 ± 1°C) and 65 ± 2% rh for 24 hours/day
over the whole year. Samples brought into the laboratory have to be conditioned for
sufficient time before testing to allow the moisture content of the textile material to
reach equilibrium with the laboratory environment. Whilst it might only take a few
minutes for a single fiber to reach equilibrium, a can full of sliver might take
several days. As a guide, a tightly wound yarn package should be conditioned for
48 hours.
In the second mode, shown in Fig. A4.1(b), sensors are fitted into the machine,
which generate signals that describe some function of the material or machine
performance; these signals are processed by a computer. The sensors should be
insensitive to changes in the environment or should have corrections applied to
neutralize any changes in anything other than the parameter being measured. The
computer may be local or it may be a central unit. The output of the computer can be
used either for control or for information or for both.
Advanced topics II: Testing of textile materials 351
Textile
processing
machine
Textile product
Sensor
Textile
processing
machine
Laboratory
Control
Computer
Testing apparatus
(a)
Data
Data
Fig. A4.1
(b)
Testing modes
A4.1.2 Laboratory testing
The priorities for testing filament yarns differ from those for staple yarns. In staple
spinning, the most common measurements relate to variations in linear density, fault
levels, fiber properties, yarn twist, and strength. Increasingly, laboratories are using
high volume instruments (HVI) to measure many fiber properties in a semi-automatic
measurement line. Other testing equipment is also becoming automated. Since the
number and type of the individual measurements are likely to vary, no description of
HVI lines will be described per se, but individual component tests will be described
under the appropriate headings. In filament testing of yarn being used for household,
apparel or industrial applications, bulk, fiber strength, elongation, and dyeability are
among the most common tests made. For household applications, cover is a very
important factor and consequently yarn bulk tops the list of priorities in these cases.
In both staples and filaments, consistency of fabric appearance is important;
consequently yarn defects, dye affinity, and variations in linear density, etc., are
becoming of increasing importance. For many industrial uses, strength is the most
important factor.
A4.1.3 Online monitoring of production
Many yarn and fiber attributes cannot be measured online because of the technical
difficulties and costs involved. However, it is now established practice for the count
of yarn, or any intermediate product, to be measured online. Also, yarn hairiness and
defect levels are commonly monitored continuously. Sensors might be connected to
devices that sound an alarm when the monitored attribute moves outside the control
limits. Any of the signals generated by the sensors involved in monitoring might be
used for control purposes, but in practice it is likely to be used only as a proxy for
mass/unit length.
A4.2
Measurements on staple fibers
A4.2.1 Fiber length
Drafting waves in roller drafting are caused when there is a significant difference
between the ratch and effective fiber lengths. The ratch settings are set to standard
values but variations in fiber length occur within any given sample, as well as from
352
Appendix 4
sample to sample. Thus, there are always problems with drafting waves and some
control of effective fiber length is desirable. It will be recalled that ‘effective fiber
length’ refers to the in situ behavior of the fiber; the effects of hooks, lack of fiber
straightness, and fiber ends are all taken into account. Thus, it is necessary to consider
not only the fiber length distribution but also how processing affects the geometry of
the fibers.
The traditional method of displaying the fiber length distribution is to progressively
comb out fibers of descending length from a prepared sample. Fibers are then placed
on a velvet board in the form of an array, as illustrated in Fig. A4.2(a) (the array
shown is for American cotton). The actual procedure needs skill and is time-consuming.
An alternative is to use a machine to create a fibrogram. The samples have to be
prepared before being measured to ensure that the fibers are parallel and that unclamped
short fibers are removed from the fringes to give a length-biased sample. This is done
by using a comber roll. A fiber fringe is produced in which one end of the beard is
clamped and the free ends extend away, with the fibers parallel to one another. A
schematic sketch of such a fringe is given in Fig. A4.2(b). Fibers are viewed in a
narrow aperture in an optical or capacitive system. The aperture is called a window
in the diagram. The signal derived from the sensor elements provides an estimate of
the number of fibers in the window. If the clamp is moved to change the dimension
X, the number of fibers viewed in the window also changes. A plot of the number of
fibers against X gives a fibrogram from which the distribution of fibers can be
deduced. The fiber array is often typified by two span length readings. The span
(a)
Fibers
Window
X
Clamp
(b)
Reaction force measured
by load cell
a and b = hooked fibers
Clamp
Clamp
b
a
Fibers
Force caused by downward
movement of lower clamp
(c)
Fig. A4.2
Fiber measurement
Advanced topics II: Testing of textile materials 353
length is the average length of those fibers that fall in the longest y% of the fiber
population. Values of y commonly used are 2.5% and 50%, the long fibers being
typified by the 2.5% span length and the short ones by the 50% span length. The span
length changes and the hooks are pulled out as the material passes through the
various draft zones. The added variance in drafting is greater for the 50% span length
fibers than the 2.5% ones.
It is normal to report the upper half mean length, uniformity index, and the
short-fiber content. Uniformity index is defined as the ratio between the 2.5% and
the 50% span lengths; it is often quoted as a percentage. The short-fiber content is
taken as the percentage of fibers less than 0.5 inches. Care should be taken in
comparing results from the various sorts of length sampling; different machines may
produce results that are not truly compatible because of differing sampling procedures.
Some testing equipment is able to report using differing sampling schemes, which is
of considerable value if test results are to be compared in different departments or
different mills.
The clamping of hooked fibers can give ambiguous results. If hooked fibers are
clamped at the base of the ‘U’ of the hook, then the fiber length, as seen by the testing
machine, is shorter than the real length. If the fiber is clamped at the extremity of the
longest leg, the measured length will be close to the real length. There are intermediate
conditions, which give a variety of errors. This is despite the combing operation,
which is part of the preparation procedure. It follows that, where there is a predominance
of leading or trailing hooks, such as with sliver, the direction in which the material is
mounted in the clamps becomes important. If there is doubt about the direction, the
material should be tested in both and the higher of the two values should be used.
Insufficient combing during fiber preparation produces error also. Some fiber breakage
will occur during specimen preparation, but the error from this source will be small
providing the fiber fringe is not too vigorously combed. A compromise between
under- and over-combing has to be found for minimum error.
A4.2.2 Fiber strength
Fiber strength is measured on a beard of fibers clamped as indicated in Fig. A4.2(c).
One clamp moves relative to the other to extend the fibers and a load cell attached to
the fixed clamp measures the load on them. If the number of fibers in the crosssection is determined, the fiber breaking load can be determined. Further if the
average fiber linear density is known, the fiber breaking stress may be calculated. On
an HVI line these measurements are made on the same samples, which helps in the
matter of accuracy of result. The system is similar in principle to that later described
for yarn. Again, hooked fibers can reduce the indicated fiber stress with respect to the
actual one; furthermore, insufficient combing will reduce the indicated stress because
of the lack of fiber orientation in the direction of loading.
A4.2.3 Fiber fineness
Fiber fineness is measured by placing a given mass of fiber in a container of given
volume and measuring the air permeability by passing air through the mass. Fiber
fineness is related to this air permeability and the results are expressed as the micronaire
index.
If a fiber is loosely packed into a standard tube and air at standard conditions of
354
Appendix 4
temperature and humidity is passed through the packed tube, a pressure drop will be
found. This air pressure drop/unit flow volume is proportional to:
k (As)2ρM2/{(Kρ) – M}3
[A4.1]
where As is the surface area of the fiber sample, ρ is the packing density of the fiber,
M is the mass; k and K are factors.
For a standard sample mass of a given fiber, the pressure drop is a measure of the
surface area of the fiber, from which the fiber diameter or linear density can be
estimated. The importance of maintaining standard conditions is evident from Equation
[A4.1]; in particular, the denominator is sensitive to error. Accuracy in sample mass
is very important. Uniformity of fiber packing in the test volume is also an important
factor and any tendency for the fibers to occlude will produce an error. Despite the
care needed to get reliable results, the test is simple and rapid and it is widely used
in mills. The result is described by the term ‘micronaire index’ (‘micronaire’ for
short). Cotton fibers from a given geographical zone often have factors k and K
which vary but little, and the test is useful within such zones. For example, the
‘micronaire’ test is widely used within the USA because of its reliability under
normal circumstances for US grown upland cotton or other varieties. However, if
cottons from many geographical zones or varieties are mixed, there could be
difficulties. Other more sophisticated instruments are available, as described in Section
A4.2.4.
In the case of flax fibers, the fibers are gummed together at the beginning of the
process and they have to be divided during processing. It is possible that some fibers
are still gummed together at the end of the process and the variability in apparent
fiber fineness might be more marked than desired. Also, since the product is a natural
fiber, the fiber fineness is variable. Nevertheless, the fineness of the fiber is a parameter
that should be controlled. The usual equipment for these measurements is a simple air
permeability tester with a short test cylinder into which a U-shaped sample is inserted.
Variations in the cross-sectional shape of the fibers affect the air permeability tests as
just described. This is a disadvantage as far as measuring fiber linear density (micronaire)
is concerned, but it brings with it the advantage of ease of use.
A4.2.4 Cotton immaturity
Immature cotton has a different cross-section from mature cotton, and the permeability
test helps identify this important parameter. Immature cotton fibers also have dyeing
characteristics that differ from those of mature fibers. Thus it is often important to
determine if the fiber being used is mature or not. The fiber fineness test described
in Section A4.2.3 is incapable of discriminating between the effects of changes in
fiber cross-sectional shape and fiber ‘diameter’ (i.e. fiber fineness). In trying to
measure two parameters with one measurement, there is always an ambiguity.
Sometimes, with a restricted source of supply, the immaturity and micronaire correlate
but outside that sphere it is necessary to conduct a second test. One way is to carry
out second permeability tests on fibers swollen by treatment with caustic soda and
make comparisons with the first test on unswollen fibers. Alternatively, double
compression tests can be used. A constant airflow device is used to measure the
pressure drop of a standard sample at two different fiber-packing densities. A second
alternative is to use image analyzers, but the expense of the equipment limits this
option. Projection and other microscopes can be used to look at single fibers or small
Advanced topics II: Testing of textile materials 355
groups of them, but the labor and equipment involved again makes this a fairly
expensive alternative.
A4.2.5 Optical character
Color and reflectance are measured by an optical system; yellowness (+b) and reflectance
(Rd) are reported. Trash content is also measured by an optical system that relies on
the fact that trash is darker than cotton. The percentage of the viewing area of the
specimen that is dark is taken as a measure of trash content. The measurements of
fiber length, strength, fineness, color, and trash content can all be made on a single,
automated machine known as a high volume instrument (HVI). The use of such
machines has virtually replaced manual cotton classing in the USA and they are
widely used elsewhere.
A4.2.6 HVI calibration
A problem that exists with the HVI relates to consistent calibration of the machines
over time and between various laboratories. Since their use is mainly with cotton,
emphasis is placed on the use of calibration cottons as a standard. The difficulty is
that even the calibration cottons are variable; consequently, statistical control techniques
have to be used to keep the machines within acceptable limits of calibration.
A4.3
Measurement of linear density of staple yarns
A4.3.1 Determination by weighing
A commonly used manner of testing for yarn count is to weigh skeins and derive an
average value. (A skein usually contains 120 yards of yarn.) However, the variations
in count from skein to skein must be closely watched in order to prevent barré in
fabrics. Commonly, this form of measurement is used both in the production facility
and the laboratory.
In a long specimen such as a skein, short wavelength errors escape detection; there
can be no information gleaned about the inch-to-inch variation in linear density.
Sliver is usually measured by a yard board (Fig. A4.3) which is simply a template
Cut
1 yd
Sliver
Cut
Hinged sliver board
Sliver
Fig. A4.3
Yard board
356
Appendix 4
for cutting a set length of sliver. Other lengths can be measured using the same
technique. It is a manual method often used in the mill, which is simple to use; it is
a test that is reasonably accurate and the procedure is fast. Again, it is necessary to
test a sufficiency of samples.
A4.3.2 Continuous measurement
The discussion in the previous section centered on the mid and upper limits of length.
now consider the lower limits. In continuous measurement, sensors are used to measure
some set of attributes that reflect the parameter to be monitored. For example, linear
density might be measured by using a pair of electrodes to measure the capacity of
the electrode gap with the staple yarn, roving or sliver passing through that gap (Fig.
A4.4). An example of this type is the Uster evenness tester [1]. A small, varying
electrical voltage across the plate enables differences in the electrical capacity of the
electrode system to be detected. As the mass of yarn in between the electrode changes,
so does the capacity. The length of the electrode, L, is only a few millimeters and so
the data stream reflects very short wavelength changes. Thus, the equipment is able
to monitor continuously the changing linear density of a running yarn and the resulting
electrical signals are converted to coefficients of variation, spectrograms, and strip
charts.
There is a relationship between the capacitance and the linear density under controlled
conditions, which makes it possible to treat the output from the monitor as a measure
of linear density [1]. As a second example, a beam of light can be used to project a
shadow or image of the passing material and to translate that image into an electrical
signal. In a properly designed system, the signal has a reasonably stable relationship
with linear density and is often regarded as a measure of it. In these and other
devices, the active sensor element is usually of the order of 1 mm, as measured along
the length of the material being measured. Consequently there is no problem in
resolving the data containing the complex spectrum. More detailed discussion of
electrode width is given by Furter [3].
Some electrical circuitry in the system as sketched in Fig. A4.4(a) has been omitted
for simplicity. Also, as a precaution against the effect of stray electrical fields, it is
normal to surround the active electrodes with guard elements. These guard elements
are set into the same plane as the active ones and are earthed, or held at a controlled
voltage. If we take each of these in turn: capacitance is controlled by length, L, the
dimensions of the gap, the cross-sectional shape of the strand, the position of the
strand within the gap, and the dielectric constant of the material in the gap. Within
limits, L can be adjusted electronically. It is important to use the correct set of
electrodes with the proper gap size for a given material. Several gap sizes (or slot
numbers) are available on the commercial testing machines. Air has only a negligible
effect, but moisture in the fibers has a powerful effect. Thus, it is important to control
the moisture content of the fibers if an accurate result is required. The effective
electrode length can be changed by integrating the results over time. Such integrated
values are used for inert tests in which the higher frequency variations are damped
out and the signal is smoothed to give a moving-average value.
One form of record (Fig. A4.4(b)) relates the deviation of linear density to time
elapsed. Since both the textile material and the chart travel at known speeds, it is a
simple matter to translate time into length of material. These charts are known
colloquially as strip charts and technically they are expressed in the time domain.
Advanced topics II: Testing of textile materials 357
L
Electrodes
Bridge
Yarn
Output
signal
Amplitude
Yarn
(a)
Amplitude
Time
Error wavelength, log scale (inches)
(b)
2
4
8
16
2
4
8
16
Error wavelength, log scale (inches)
(c)
Fig. A4.4
Capacitive transducer system
Strip charts have their main value in making long-term variations visible. In this
regard, long term relates to the length of textile material tested. Extra long-term
trends cannot be seen. The charts are also valuable for detecting irregular yarn faults
and disturbances that, because of their non-repetitive nature, do not show up on a
spectrogram.
A second form of diagram is the spectrogram, which is a chart expressed in the
frequency domain where repeating patterns of deviations are resolved by frequency
or wavelength rather than time (Fig. A4.4(c)). For example, a simple sine wave can
358
Appendix 4
be expressed as a single line at the appropriate frequency or wavelength in the
frequency domain. A complex wave made up of sine waves gives a spectrum of lines,
but the rendering of the data is more economical and understandable in the frequency
domain.
To be effective, a sufficient length of textile material has to be tested and a large
amount of raw data is generated. As implied in the previous paragraph, a way of
compressing data is to convert what, in essence, is a long time series into the frequency
domain. To further compress the error wavelength scale, it is usual to use a logarithmic
basis. In the terminology of textile processing, the chart expressing evenness data in
the frequency domain is known as a spectrogram although, technically, it is a
periodogram. The spectrograms in Fig. A4.4(c) show reasonably good and bad examples
of roving evenness. The first is an example of a spectrogram of a very bad yarn,
which shows two humps (probably caused by the interaction of a bad front roll on a
roving frame, combined with improper roll settings). The second shows a spectrogram
that would often be regarded as satisfactory. In the spectrograms, the ordinate is often
merely referred to as amplitude, but this needs a little more discussion. The purpose
of such equipment is to record mass variations or so-called ‘evenness’, ‘regularity’,
or the negatives of these. Based on the work of Furter [4] and others, one may define
the parameter in two ways. The first is:
Mean deviation or U% = [100/(xmT)]
∫
T
| x i – x m | dt
[A4.2]
0
where xi is the instantaneous value of linear density, xm is the mean value, | xi – xm |
is the deviation of the instantaneous value from the mean, T is the evaluation time,
and t is time in compatible units.
If s is the standard deviation, s2 = ∑ (xi – xm)2/(ns – 1), and CV = s/xm, the second
definition is:
CV % = [100/ x m ]
√
(1/T )
∫
T
0
( x i – x m ) 2 dt
[A4.3]
Subject to the data being Gaussian or normally distributed, the relationship between
U% and CV% is stated to be CV = 1.25 × U%. Furter discusses in detail the meanings
of various strip charts and spectrograms, as well as the measurement and importance
of yarn faults.
Turning to online monitoring, suppose a device is measuring, online, the linear
density of sliver emerging from a draw frame at 600 m/min, and that the output
signals relate to successive 1 mm lengths. Each measurement requires a calculation
and there would have to be 1000 × 600/60 = 10 000 calculations/sec. A central
processing unit in a computer adequate for this traffic thus becomes essential. In a
24 hr day, there would be 864 million calculations for each measurement position.
There has to be some means of filtering the output, otherwise the analyst would be
overwhelmed. One way of filtering is to use periodic measurement, but if the
measurements are spaced too far apart there might be difficulties with under-sampling.
Another way consists of the equivalent of a control chart and provides a warning only
when the parameter concerned moves outside the control limits.
An example of a control chart is given in Fig. A4.5, in which three values with
round shaded plot symbols represent out-of-control points or outliers. Only these
three points would be reported and the rest would be ignored except, perhaps, for a
Advanced topics II: Testing of textile materials 359
6
Upper control limit
5
Trend
Value
4
3
Lower control limit
2
1
0
0
5
10
Time (arbitrary units)
Fig. A4.5
15
Control chart
trend analysis that, in this case, shows the variables going steadily towards the upper
control limit. As an aside, it should be noted that a record may be kept in the
computer memory only for a limited period. When the data generated from the
transducer is used in a control system, there are a number of ways to conserve
computing power and keep the system under stable control. A transducer is a device
that converts the variable to a usable signal. The digital data stream can be compressed
to facilitate the transmission of signals to the controllers. Signals can be restricted to
outliers, trend analysis can be used to modify the control, and other schemes can be
applied to keep the volume of transmitted data within bounds.
A4.4
Measurement of twist
A4.4.1 Untwisting to zero twist method
With ring and twisted filament yarns it is possible to untwist yarns until the fibers are
approximately parallel, and this condition can be determined with a fair degree of
accuracy.
With filament yarns, the condition of zero twist can be determined by placing a
thin blade between the filaments and then sliding the blade along the length of the
yarn within the gage length [2]. When the blade can be moved without resistance
from one end to the other, the yarn is at the zero twist condition. However, the use of
such techniques becomes difficult with staple yarns. Ring yarns pass through what is
essentially zero strength as they are untwisted, to zero twist. However, this is not true
for rotor yarns. Zero twist staple yarns have little or no strength unless a bonding
agent is present (which is highly unusual at this stage of processing). Consequently,
the gage length is normally set below the staple length so that a sufficiency of fibers
is gripped at both ends. It is desirable to test yarns under some tension to prevent the
yarn from snarling when twisted. The tension applied when the yarn is at or near zero
twist should not be so great as to cause the weakened yarn to break. The apparatus for
such a test consists of two clamps attached to the yarn and some means of creating
a sufficient controlled tension in the yarn between the clamps. One of the clamps
rotates and untwists the yarn and the other is fixed.
The test is fairly labor intensive and is usually carried out in the laboratory.
360
Appendix 4
A4.4.2 Twist contraction in staple yarns
Twist is measured in turns per unit length and, as the twist is changed, the yarn
changes length. The change in length from the untwisted to twisted condition is
known as twist contraction and it can be used to help measure twist. This should not
be confused with the contractions that take place when filament yarns are textured.
A4.4.3 Reversed twist method used in staple yarn testing
A second technique requires one to carry on with what was the untwisting into the
zone where the yarn becomes reverse-twisted [5]. Reverse-twisting continues until
the gage length regains its original value, it being assumed that the twist contraction
in the reverse direction is the same as the original direction. The reversed twist
method is often known as the twist/untwist method (even though it would be more
accurate to say ‘untwist/twist’ method). As shown in Fig. A4.6, a counterweight
controls the yarn tension and the twist counter reads the change in the number of
turns from the beginning to the end of the test. The twist indicated is assumed to be
twice the twist in the yarn; consequently the twist density is half the indicated twist
change/gage length. The initial tension in the twisted yarn should be maintained at
0.25 + 0.05 g/tex (or 2.45 + 0.49 mN/tex). The tension at the end of the reversetwisting should be the same as the initial value.
A4.4.4 Twist testing rotor spun yarns
Rotor spun yarn never achieves a state during untwisting where all the fibers are
approximately parallel. Thus, although the reversed twist method just described is
normally used, there is an error due to the structure of the yarn. The measured values
differ from those calculated from the machine parameters, as shown in Fig. A4.7 (for
a polyester/cotton blend in this case). In this diagram, there should have been no
change as the percentage polyester was altered. In the case of the so-called machine
value calculated from the known rotational speeds of the machine components, this
was true. The machine value was then used to normalize the other results. With the
reversed twist method, changes in blend did affect the measured twist and this implies
differences in yarn structure. This error is normally ignored because a standard of
Hinged clamp with counterweight
Twist counter
Yarn
Rotating clamp with motor
Fig. A4.6
Twist tester
Advanced topics II: Testing of textile materials 361
Twist testing factor
1.0
Machine
Reversed twist
0.5
Surface fiber
12/1 Polyester/Cotton OE yarn
0
Fig. A4.7
0
50
100
% Polyester in the blend
Various methods of twist measurements for rotor yarns
judgment different from that used with ring yarns is applied. The effect varies with
the fiber being used. If one judges the twist by the fibers on the surface of the yarn,
there is even more error, and so this method is little used. The population of bridging
fibers in rotor spun yarn usually varies from 10 to 25% and most of these produce the
wrapper fibers on the yarn surface.
The fibers behave as if they are shorter than they really are and, as mentioned in
Section 7.2.12, a higher TM is required than would be used in ring spinning.
A4.5
Visual examination of yarns
A4.5.1
Yarn board
One factor in assessing staple yarns is that of yarn grade. Fairly long samples of the
subject yarn (which are normally white) are wrapped on a black yarn board (Fig.
A4.8) or sleeves are knitted from it. These techniques are used to telescope the errors
and make them more visible; they are useful for determining fault rates, short and
mid-term errors of both staple and filament yarns. However, it cannot show errors
longer than a fraction of the sample length, which is about 100 yards or so. A yarn
blackboard is either a rectangular or trapezoidal board onto which yarn is wound,
closely spaced to simulate fabric. Periodic errors produce patterning to appear on the
yarn blackboard, which indicates a mechanical error. Slubbiness indicates drafting or
drawing defects, and reflective differences indicate changes in yarn luster, or yarn
Yarn
Slot to hold yarn end
Fig. A4.8
Yarn board
362
Appendix 4
hairiness, or both. For dyed filament yarns, differences in color, luster, and bulk can
be assessed. The eye quickly becomes skilled in assessing the error characteristics
and in judging the quality of the yarn. A yarn board is a very successful and simple
device. However, it cannot show errors longer than about 100 yards or so because
sample length is insufficient. The example of use in checking for variation in luster
and/or bulk in a textured yarn has already been mentioned. Textured yarn has to be
wound at constant tension because the structure of the yarn is such that the bulk
changes with tension; thus surface reflectivity and yarn diameter also change. A
knitted sleeve can be used to achieve a similar objective. In this case, the knitted
material is robust enough to be dyed and, as explained in earlier discussions, this
dyeing may well reveal faults created in previous processes. Faults caused by different
temperatures or mechanical stresses would otherwise not be visible until it is too late
unless differences in dye affinity can be harnessed in the test procedure. This is why
the dyed knitted sleeve is valuable.
A4.5.2 Electronic yarn boards
Electronic equivalents to the yarn board have been developed (e.g. CyrosTM) in which
evenness data are used to produce a raster.1 The brightness or thickness of the line
being painted on the computer screen or paper is proportional to the diameter of the
element of yarn being portrayed. The data can be acquired at the yarn testing stage or
by online sampling using capacitive or optical sensors. In the latter case, no yarn
would be consumed for testing. Each long series of data is divided into sub-series by
deliberate choice and the selected series are plotted in the x and y directions on a
monitor or print-out, to simulate a woven fabric. Such equipment is beginning to be
used for fabric simulation and quality control in yarn manufacture. To be able to
project what a given yarn will be like in fabric form is a useful addition to the means
available to be able to control quality and assist sales. Furthermore, it can be done
without the expense of actually making the fabric. It is likely that such an arrangement
will permit recognition of very long-wavelength faults that are often missed with the
present technology.
A4.6
Yarn hairiness in staple yarns
For the purpose of discussion, a fiber projecting from the surface of a staple yarn will
be called a hair. It is, of course, sometimes difficult to define the surface of the yarn
and the definition of hair is not very precise either. Error in measurement is thus
inherent in some methods but such errors are better than no measurement or control.
There are a large number of techniques [6] for measuring hairiness which range from
microscopy to online measurement of projected optical or electronic images. Error in
measurement is inherent in some methods but such errors are better than no measurement
or control. For example, referring to hair A in the plan view of Fig. A4.9, the projecting
hair length, when viewed in elevation, is foreshortened as shown. In addition, part of
the hair length may be obscured by the body of the yarn, as also is shown. To increase
1 A common example of a raster is a television set where a light beam is made to oscillate across
a screen in a two-dimensional pattern and the brightness is modulated to produce a picture.
Advanced topics II: Testing of textile materials 363
Elevation
W
Projected
length
A
Plan view
Fig. A4.9
Yarn hairiness
sensitivity of measurement, a mask of width W is often added; only the variations
outside the shaded area are measured. If the width is smaller than the yarn diameter,
signals from the body of the yarn are added to the signal as shown by the black areas
bounding the shaded rectangle. This too produces an error. If W is too large, another
error is introduced. Since yarn diameter is variable, some error is inevitable if such
masks are used, but the increase in sensitivity reduces other errors in measurement
and the net result is that it is worth using them. Often the information required relates
to the outer surface of hairs, and the masking width is then increased to eliminate the
portions of hair close to the yarn body. In these cases, the mask width, W, is a multiple
of the yarn core diameter. Hairiness is variously described by the number of hairs/
unit length at a given radius, or by the length of the hairs. The hairiness is dependent
on the yarn diameter and has a roughly linear relationship with fiber length and twist
over a reasonable interval.
Many different types of hairiness measuring equipment are used in laboratories
and online means of testing are available. Outputs give hairiness data in both the time
and frequency (wavelength) domains. Spectrograms of hairiness are becoming as
common as spectrograms of linear density. Much of the comment made earlier about
variability of linear density also applies to variability in yarn hairiness.
364
Appendix 4
A4.7
Tensile testing of strands
A4.7.1 Testing single strands
A tensile testing machine consists of two yarn clamps that grip the yarn specimen and
move apart to induce load in the specimen. Usually one clamp is held stationary and
the other moves. The stationary one usually holds the transducer or load cell, which
measures the load applied, and the movement of the other clamp reacts with a sensor
to give the elongation of the specimen. Electrical or other circuits control the elongation
of the specimen by changing the movement of the moveable clamp. The design of the
load cell is an important part of the specification of the machine. Most load cells
require a finite elongation in the direction of load to produce a signal. It is highly
desirable that the system should be stiff so that the deflection of the machine is very
small in comparison with the strain in the specimen. A discussion of the design of
transducer is given by Furter [7].
Care has to be taken in interpreting the results because of the length factor. A long
specimen has a low strength because of the number of weak spots in its length,
whereas a short sample has more variable results. The short-term coefficient of
variation of tenacity is of considerable importance in determining the efficiency of
the ring spinning and subsequent operations. Despite the arguments for using singleend yarn testing, it is more expensive than skein testing and involves high skill by the
technician. For these reasons, one finds mostly skein testing in the mills and singleend testing in research laboratories. Also, the materials are visco-elastic, which means
that the results are dependent on the rate of loading of the material. This becomes an
important matter when results from different test facilities have to be compared.
A sample should be tested in a representative fashion. One implication of this is
that the correct length of specimen should be tested. The tests often yield the breaking
strength and the elongation at break. Because most textile fibers and yarns are viscoelastic, the stress–strain curve is not linear. Consequently, if one wishes to characterize
the behavior in normal use, where the loading is not high, attention has to be paid to
the slope of the stress–strain curve near zero load. Examination of elongation at
break gives data that helps one assess the visco-elastic nature of the yarn or fiber. For
complete information about the visco-elastic behavior, it is necessary to cycle the
load to determine the hysteresis, but such testing rarely enters into commercial mill
practice.
Tensile test machines fall within one of three categories [8], namely:
1
2
3
Constant rate of extension (CRE);
Constant rate of traverse (CRT); or
Constant rate of load (CRL).
The test category should be cited with the test results because the type of test affects
the numerical results. Selection of the category depends on the end use of the product
being tested. It follows that inter-laboratory comparisons must be for like categories.
In strength testing, the values are nearly always measured offline because the test
is destructive. The shortest effective sample in measuring strength is where the yarn
clamps touch one another at the clamping points and the gage length is zero (Fig.
A4.10(a).
Two problems arise. The first is that the yarn clamps induce a stress in the yarn
around the clamping zones. This influences the characteristics of the yarn actually
tested, and usually yields a lower breaking force than should have been obtained. The
Advanced topics II: Testing of textile materials 365
Clamp
Clamp
Gage length
Gage length
Clamp
Yarn
Yarn
Variations in yarn
thickness are
exaggerated
Clamp
(a)
(b)
Fig. A4.10
Tensile testing of yarn
second problem is related to weak link theory and yarn variance. A chain always
breaks at its weakest link and a yarn sample can be regarded as a chain of infinitesimally
short links. A zero gage length test produces a variance amongst the results similar
to that existing in the length of yarn from which the test specimens were taken. A
long gage length contains a distribution of link strengths (Fig. A4.10(b)) and the
weakest one fails; an erroneous conclusion would be that the single result is typical
of the whole length tested. With insufficient testing to establish the variance, the
single result would be no more than an approximation without any knowledge of the
probable range of error. Clearly a long gage length has a greater probability of
including a weak link and producing a low result than does a short gage length for
that single test. Often a long sample is wrongly preferred because it is thought that
the results are less variable and, therefore, the results are more reliable.
There is a wide range of stress–strain characteristics of fibers and yarns. Space
precludes more than a sample, but Fig. A4.11 shows a wide range, especially between
untextured filament and staple yarns. It will be noted that there are also wide differences
in the shapes of the curves that reflect their visco-elastic character. It must be emphasized
that the curves shown are single-end tests and that there are considerable variations
within any one series of tests with a given yarn.
366
Appendix 4
500
Nylon filament
Polyester filament
Stress (mN/tex)
400
300
Cotton yarn
200
Polyester staple yarn
100
0
0
Fig. A4.11
10
Elongation (%)
20
Stress–strain curves
A4.7.2 Testing skeins
With the testing of single yarns as sketched in Fig. A4.11, the results should be
averaged after sufficient tests have been made. An alternative method to obtain a
value for yarn strength is skein testing [9]. However, if a skein is tested, many parallel
portions of yarn are loaded simultaneously. Since all fibers gripped by both clamps
experience the same extension, the load suffered by each gripped fiber depends on its
longitudinal stiffness. A stiff fiber may take more load than less stiff parallel portions,
with a result that certain fibers will fail before others have reached their breaking
stress. Thus, the skein may appear weaker than might be expected. On the other hand
there is an averaging effect between the parallel portions which reduces the effects of
weak spots in the yarn. Long specimens of yarn in the form of skeins give strength
values that are some sort of average; because of that, they are frequently used. Skeins
of 120 yd are commonly used.
The strengths of both single yarns and skeins are often expressed in terms of
tenacity. For a single yarn, the tenacity (T ) often employs the units mN/tex. For a
skein, the mathematical product of cotton count and breaking load in lbf (CSP –
count-strength product) is customarily used.
ASTM D1578 suggests that:
CSP = KT
However, varying characteristics of yarns and skeins cause the factor K to alter from
case to case. ASTM quotes K = 21.23 for cotton.
Advanced topics II: Testing of textile materials 367
Force
Skein
Force
Fig. A4.12
Skein test
Variation from skein to skein gives no information about short-term errors because
these have been averaged out, but does give information about the earlier processing
stages. Practice varies somewhat but an example typical in the USA is that 20, 40, or
80 turns of yarn are wrapped on a reel of 1.5 yd circumference to form the skein. The
skein is usually elongated on a tensile testing machine at 12 inches/min.
A sketch of a skein mounting portion of a testing apparatus is given as Fig. A4.12.
368
Appendix 4
The result is quoted as the lea strength or CSP. The skein test is considered unsuitable
for yarns that stretch more than 5% during the test. Some low cost testing machines
have a pendulum arm fitted with a ratchet. When the moveable crosshead drops, the
load on the pendulum lever causes it to rise along a quadrant on to which the ratchet
mechanism is fixed. The weight on the pendulum rises with load, but when the skein
breaks it is unable to drop back and so it records the breaking load.
A4.7.3 Multi-function fiber measurement
Because of the number of tests needed to control the blending of natural fibers,
attention has been turned to assembly line layouts reminiscent of the automotive
industry. This is particularly so with cotton testing. The so-called high volume
instruments (HVI) consist of several working stations situated along a console. At
each working station, one or more fiber attributes are measured that differ from the
attributes measured elsewhere in the HVI. Data from every transducer is stored and
processed in a computer. Thus, the testing of a sample is carried out at virtually the
same time, under the same atmospheric conditions, with the sub-samples needed to
make fiber beards taken from zones in close proximity. This arrangement expedites
the flow of work and reduces the probability of error. Current HVI machines measure
trash content, short fiber concentration, fiber fineness (micronaire), upper half mean
length, strength, elongation, color, and reflectance. The acronyms used are trash,
SFC, MIC, UHL, STR, ELO, +b and Rd. An earlier discussion (Section 11.2.1) deals
with the definitions and importance from a quality control standpoint, whereas some
other attributes will be measured in the future, and there certainly will be developments
in the techniques of measurement.
The Advanced Fiber Information System (AFIS) device measures fiber properties
while the fibers are being carried by an airstream. A scanning laser illuminates the
moving fibers and transducers pick up the reflections. The fiber fineness, length, and
color attributes can be assessed in this manner. The device is not in such wide use as
the HVI and it does not measure fiber strength and elongation.
A4.8
Filament yarns
A4.8.1 Linear density of filament yarns
Normal untextured filament yarn has a linear density, which is virtually invariable
along its length. Consequently, it is sufficient to reel off a certain length, and weigh
it under standard conditions of temperature and relative humidity of the workspace.
However, when the yarn is textured, the mass may no longer be evenly distributed
along the length. Also, the modulus of elasticity of the yarn is low due to the texturing
(i.e. it is ‘stretchy’). Thus, textured yarns have to be tested under constant tension
conditions. Because of the variability, methods similar to those explained for staple
yarns may be employed.
A4.8.2 Strength of filament yarns
For industrial uses, filament yarns are usually tested for strength. This is a straightforward
tensile test; providing there has been proper sampling, it is a simple exercise in
Advanced topics II: Testing of textile materials 369
quality control. For apparel yarns, tensile tests are also used on occasion. The purpose
in this case is usually to check for degradation of the filaments by overheating in
texturing, or by some other cause. The tenacity loss can be up to 10% with polyester
and nylon. Proxies for poor quality can be found by observing the manufacturing
equipment. Excessive ‘snow’ (powdered oligomer and finish) around a texturing unit
and excessive production of fumes from a heater are two such examples. Also, tests
of evenness can be used to show errors. The problem is sometimes in the quenching
after extrusion. Typically, a false twist textured nylon or polyester has a tenacity of
350 to 500 mN/tex (4 to 6 gf/den) and an elongation at break of between 25% and
35%.
A4.8.3 Definitions relating to the bulk of textured yarns
There are a number of terms concerning normal usage of textured thermoplastic
filaments that need to be mentioned. Bulk is generated when filaments are caused to
coil, to take up a zigzag shape, or to be deformed in any micro-convolution. Some
early forms of texturing involved a zigzag texture, and bulk was then described in
terms of crimp. This terminology has been extended by use to cover other sorts of
texture. Thus, the ability for a yarn to contract under tension is called crimp contraction
and the ability to recover is called crimp recovery. Bulk shrinkage is a term relating
to the potential stretch and ‘power’ of stretch yarns, or a measure of bulk in textured
yarns [10]. (The term ‘power’ is used to convey the recovery properties of so-called
elastic yarns and fabrics; the term ‘elastic’ means the ability of the material to
withstand large deformations, and does not relate to the engineer’s definition as the
relationship between stress and strain.) The term ‘crimp’ is used as a proxy for bulk,
which is hard to define. A rough equivalent to bulk is yarn diameter, but the diameter
is transient in the sense that it changes under load; also, when assembled into fabric,
the cross-sectional shape changes according to the loads applied by the intersecting
yarns.
ASTM standard D4031 [10] defines crimp contraction, in this context, as ‘an
indicator of crimp capacity or a characterization of a yarn’s ability to contract under
tension’. Crimp recovery is defined as ‘a measure of the ability of a yarn to return to
its original crimped state after being subjected to tension’.
When a textured yarn develops bulk, it shrinks, even under load. Some yarns have
a structure that favors ease of elongation (so-called ‘stretch yarns’) and others favor
yarn bulk. The elongational behavior and the hysteresis loss (ability to recover from
deformation) can be measured on tensile testing machines, as described earlier.
A4.8.4 Measurement of bulk and crimp in textured yarns
Fabrics made from bulked yarns are intended to provide cover and insulation. The
bulking processes are varied in nature and produce a wide range of yarn structures.
Also the polymers available cover a wide range. Because the range of performance
varies widely, there are several tests and sets of conditions available. ASTM suggests
several loading options, which may be summarized as (a) 0.04 to 0.98 mN/tex, and
(b) 8.8 mN/tex, the first being sufficient to extend the yarn without removing crimp
and the second to remove crimp without significantly elongating the filaments. The
recommended protocol prescribes that the low loading should be kept in place during
the test, and that an extra load be added and removed when necessary to adjust the
370
Appendix 4
load between the two levels. Also, bearing in mind that the materials are visco-elastic,
times of heating and loadings are strictly detailed as are other aspects of this series
of tests. Information is kept up to date by ASTM, to whom the reader is referred for
more information. During tensioning of the yarn under test, it can be immersed in
water (Fig. A4.13), subjected to dry oven heat, or steamed. It is recommended that
textured polyester yarns be tested first at the low load condition and then at the higher
level of load. With textured nylon or polyester yarns, the water bath method can be
used. For polyester yarns the bath should be at 97°C and a low stress level used,
whereas for nylon a temperature of 82°C and a stress level of 0.13 mN/tex is
recommended. The skein size is usually determined from the reel diameter (usually
1m) and the linear density of the yarn. ASTM standard D4031 quotes the numbers of
turns on the reel varying between 25 and 63, according to the linear density of the
yarn.
Crimp contraction is quoted as the percentage change in length between the high
and low load conditions. Similar measurements are made before and after heating to
develop crimp. Let the length of the skein at low load be X and at high load be Y. Also,
let the condition before heating be designated subscript 1, and after heating by subscript
2. Thus, the length before heating at low load would be written X1, etc. After heating,
(a) the skein length at low load is measured, (b) the extra loading is applied to bring
Water-filled
glass
measuring
cylinder
Skein
Weight
Fig. A4.13
Yarn bulk
Advanced topics II: Testing of textile materials 371
up the load to the high level, (c) the skein length is measured again, (d) the extra
loading is removed, and (e) the skein length is measured for the last time. Let X3 be
the length after heating and removal of the heavy load in the stage (d) just mentioned.
1
2
3
4
5
Skein shrinkage before heating
Skein shrinkage after heating
Skein shrinkage
Bulk shrinkage
Crimp recovery
= 100 (Y1 – X1)/Y1
= 100 (Y2 – X2)/Y2
= 100 (Y1 – Y2)/Y1
= 100 (X1 – X2)/X1
= 100 (Y2 – X3)/(Y2 – X2)
Items (1) and (2) give indications of how much mechanical and molecular forces
play in the total relaxation and this information is useful for diagnostic purposes.
Items (3) and (4) give the total shrinkage at high and low loads respectively. The
crimp recovery, item (5), is the difference in length of the skein after heating caused
by the final removal of the heavy load, which characterizes the hysteresis in the
system.
A standard test for bulk is to use a modified skein shrinkage test (Fig. A4.13). The
main difference between (a) the shrinkage test as a measure of potential stretch, and
(b) the test as a measure of bulk, lies in the applied load. Care has to be taken when
measuring bulk to ensure that all samples are tested with the same degree of ‘lag’. A
freshly textured yarn behaves differently from a yarn that has stood for an hour or so;
this is because of stress decay in the thermoplastic material. Fresh yarns generally
have higher skein shrinkage and lower strength than aged ones. The properties of the
textured yarns can continue to change even over a period of 100 hours, but the change
rate diminishes with time and eventually stabilizes. The highest change rate is in the
first hour. In fabric form, finishing and other processing can cause further bulk to be
generated and this is usually associated with perceptible shrinkage.
An alternative method of measuring bulk is to measure the volume taken up by a
piece of fabric of a known mass, M grams. If the thickness is t mm and the area is A
m2, then the bulk, 1/ρ, can be quoted in m3/g, and:
1/ρ = tA/1000M
[A4.4]
The structure of the fabric affects the cover factor and ‘basis weight’ (mass/unit area
of fabric); for comparative purposes, the test is quite useful. The thickness is measured
by a standard fabric thickness tester, which compresses the fabric slightly during the
measurement (therefore there is some error).
A4.8.5 Stretch yarns
Stretch yarns fulfill a function different from that of bulked yarns. The objective of
using a bulked yarn is to cover and insulate. A stretch yarn is designed to permit
extraordinary extensions in fabrics made therefrom, thus the high load described in
the previous section may not fully extend the yarn, and different levels may be
necessary for testing some yarns. However, the procedures are similar to those described.
For stretch yarns, some use a standard weight of 20 grams acting on a skein of
12 500 denier, to give a specific stress of 0.141 mN/tex, whereas with a bulked yarn,
a weight of 2 grams is used to give a specific stress of 10% of that just quoted. These
figures are quoted merely to emphasize the difference between stretch and bulked
yarns.
372
A4.9
Appendix 4
Visual tests of fabrics
The simplest form of dye test is to knit a sleeve, dye it, and then make a visual
assessment. There are standard procedures for the dyeing of the samples. The most
sophisticated form of this sort of test is to dye a long knitted sleeve and subject it to
automatic color testing. In this, the color is analyzed for its tri-stimulus components.
The data can be recorded by a computer and related to fixed standards. Analysis of
the averages and variances permits diagnostic work to be performed, and this, in turn,
permits good control of quality. The test fabrics are usually inspected visually for
filamentation, tight spots, etc., at the time the color tests are performed.
Continuous fabric inspection systems are useful for error diagnosis. Equipment
using two scanning lasers may be used; a computer is programmed to recognize
various types of faults and the data can be used to improve the quality of the yarn. Of
course, it is far too late to wait until the fabric has been made to measure yarn faults.
Nevertheless, there is hope that the techniques can be exploited to examine the yarn
directly at an earlier stage.
References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
Anon, Uster News Bulletin, No 35, Zellweger Uster AG, Uster, Switzerland, 1988.
D1425, Annual Book of ASTM Standards, Section 7, ASTM, Philadelphia, USA.
Furter, R. Manual of Textile Technology, Evenness Testing in Yarn Production: Part II, The
Textile Institute, Manchester, UK, 1982.
Furter, R. Manual of Textile Technology, Evenness Testing in Yarn Production: Part I, The
Textile Institute, Manchester, UK, 1982.
D1422, Annual Book of ASTM Standards, Section 7, ASTM, Philadelphia, USA.
Barella, A. Yarn Hairiness, Text Prog, 13, 1, 1983.
Furter, R. Strength and Elongation Testing of Single and Ply Yarns, Mannual of Textile
Technology. The Textile Institute, Manchester, UK, 1985.
D76, Annual Book of ASTM Standards, Section 7, ASTM, Philadelphia, USA.
D1578, Annual Book of ASTM Standards, Section 7, ASTM, Philadelphia, USA.
D4031, Annual Book of ASTM Standards, Section 7, ASTM, Philadelphia, USA.
Appendix 5
Advanced topics III: Staple yarn structures
A5.1
Theoretical yarn structures
DeWitt Smith [1] stated that the basic geometrical features of a yarn determine the
resolution of fiber tensions into components parallel and normal to the yarn axis. The
summation of the components parallel to the axis provides the yarn strength and the
normal components produce compressive forces that provide frictional cohesion. If a
bundle is made up of parallel fibers and then twisted, it produces a helical structure
somewhat similar to that shown in Fig. A5.1(a). The fibers are under tensions TL and
TR, which vary according to the load applied. Interfiber friction can cause dissimilarity
between TL and TR. The component tensions acting on the fiber at the various positions
along it produce resultants directed towards the center because the fiber is wrapped
around the core of the yarn. The radial compression arising from reactions of the
many fibers, like the one shown in Fig. A5.1(b), causes each layer to compact the
layers underneath. This increases the frictional restraints acting on the fibers in the
core. Staple fibers could not survive in such a structure because the surface would
abrade leaving the next layer vulnerable. This, in turn, would abrade and the whole
structure would fail in short order. In fact, the staple fibers migrate radially during
processing so that a single fiber occupies many different radial positions along its
length. This phenomenon is known as (lateral) fiber migration; it causes the structure
to interlock so that it retains its integrity over a surprisingly wide range of conditions.
A simple experiment can be made with a piece of string. Helically wrap the string
round a person’s bare arm, and apply moderate tension. The subject will feel the
inward pressure referred to and the string will be seen to bite into the flesh. In Fig.
A5.1(c), the same situation is depicted with a fiber. If the yarn is sliced along the
plane XX, the view normal to that plane in direction Z is as shown in Fig. A5.1(d).
For a small angle of wrap, the fiber may be considered to lie in the plane of the cut.
Thus we consider the tensions T1 and T2 rather than T3 and T4. A section of the core,
in the plane XX, is elliptical as shown. Portion AB of the yarn projects as A′B′ in the
lower part of the diagram; this is the small arc of contact between the fiber and yarn
considered here. The length AB measured along the yarn is δL. The fiber lies at a
374
Appendix 5
(a)
Pressure between fiber and core
acts towards the yarn axis
p
TL
F
F
TR
p
(b)
Z
X
T3
T4
X′
dθ
A′
X
B′
T1
(c)
dθ
Fy
(d)
Fig. A5.1
X′
T2
Fibers in yarn
radius r with respect to the center line of the core. The twist is τ tpi and the force/unit
length is:
FR /δL ≈ K(T1 + T2)/R
[A5.1]
where R is the radius of curvature of an ellipse which is a function of (r, τ); K is a
factor; and δL ≈ 2R δθ.1
Equation (A5.1) shows that the compressive force is a function of fiber tension,
yarn diameter (or count), and twist level. The greater the twist, the smaller is the
radius of curvature of the yarn surface, and the larger is the compressive force. Thus,
the higher the twist, the tighter (or leaner) is the yarn.
1 Radius of curvature = [d 2 y/dx2]/ζ where ζ = {1 + (dy/dx2}3/2. The equation of an ellipse is p2
= ax 2 + by2, differentiating w.r.t x, 0 = 2ax + 2by (dy/dx) and differentiating again 0 = 2a +
2b[y(d2y/dx2) + (dy/dx)2]. When dy/dx = 0, d 2y/dx 2 = – a /by and R = –by/a. The ratio of the
major and minor axes is b/a and b/a = 1/sin α, y = r, |R| = r/sinα and the helix angle α is
controlled by the twist.
Advanced topics III: Staple yarn structures
A5.2
375
Actual yarn structures
A5.2.1 Fiber migration in ring spinning
Consider fibers traveling in the direction shown within the zone afdeh in Fig. A5.2(a).
Fibers similar to the one marked abcd are typical of migrated fibers inherited from
the roving, which pass into the twist triangle, def. Other fibers have lesser amplitudes
of lateral displacement such as the fiber marked jg. Fibers along the nip line are
squashed by the front drafting rolls such that it is very difficult for the fibers to slip
with respect to the rolls. The strand is then translated into a roughly circular crosssection from the point d downwards by the application of torque. Fibers in or near the
selvages of the twist triangle, such as the one marked de and df, take up much of the
load created by the departing yarn and become quite highly tensioned. Many of the
central fibers, such as that marked dg, cannot be taken up as quickly as they are
delivered; they bear little or no load, and they buckle. In buckling, they tend to move
to the outside of the newly formed yarn. Fibers passing down near the selvage of the
twist triangle have the highest tension and tend to migrate into the center of the
forming yarn structure, to relieve some of the tension.
Twisted roving is usually used and the roving twist provides a supply of fibers that
periodically change their lateral positions across the ribbon of fibers approaching the
triangle in a roughly sinusoidal manner. The portion abc of the yarn represents part
of one of the fiber sinusoids. Also we consider fiber abc because it is one with a
maximum lateral displacement amplitude and will, therefore, show the periodic effect
most clearly. The parts are marked a′, b′, and c′ in Fig. A5.2(b) at a later time.
(Meanwhile the portion in the vicinity of c that was highly tensioned now passes into
the center of the yarn near d.) When the portion b′ of the same fiber eventually passes
into the yarn, it is likely to be slack and to migrate to the outside of the newly made
yarn. Comparing Figures A5.2(a) and (b), it will be realized that a single sinusoidal
fiber moves across the nip line as it flows towards d. Even if the lateral displacement
is not sinusoidal, lateral movement will still occur and produce a similar effect. Thus,
some parts of a fiber are taut as they pass through the twist triangle and some parts
are slack. Consequently, the fiber migration is periodic. Changing the roving twist
can alter the characteristics of the yarn. The result of this is that the structure interlocks
into a stable structure. The phenomenon is known as (lateral) fiber migration. A
typical result is shown for a single fiber in Fig. A5.2(c). Imagine many of such fibers
in a yarn. It becomes clear that the structure loses some of its order and that fibers
pass from layer to layer, causing the structure to interlock as just mentioned. The loss
of order results in more volume being required to accommodate the fibers; in other
words, the yarn becomes more bulky. The enhanced insulation properties given by
the extra airspaces in the yarn give fabrics made from the yarn a warmer, softer feel.
Also, the fibers have room to deflect more easily, which improves the hand.
A5.2.2 Yarn hairiness in ring yarn
A good reference in the matter of yarn hairiness is given by Barella [2]. The geometry
of the twist triangle not only controls fiber migration but also helps determine hairiness.
Some of the buckled fibers extend from the surface of the yarn as hairs and loops. Air
currents in the exit nip of the rolls also affect the process of creating hairiness,
although this is more important in high speed spinning systems like air-jet spinning.
Further hairiness is created in ring spinning as the yarn passes through the traveler,
376
Appendix 5
Roving input
a
j
h
b
f
x
c
g e
Nip line
Taut
d
Taut Slack
Yarn output
(a)
a′
b′
Nip line
Taut
c′
Taut Slack
Yarn output
(b)
Typical
migrated
fiber
(c)
Fig. A5.2
Fiber migration at the twist triangle
Advanced topics III: Staple yarn structures
377
and again when the yarn is rewound onto cones or cheeses. A photomicrograph of
portions of ring yarn showing the structure and hairiness is given in Fig. A5.3. There
are single hairs and loops standing out from the surface. (It might be noted that the
lighting of the yarns shown was adjusted to show each fiber structure most clearly,
and, in consequence, the backgrounds vary.)
(a)
Ring yarn
(b)
W
W
W
W
Air-jet yarn
(c)
W
Rotor yarn
(d)
Ring yarn
(e)
ST yarn
Fig. A5.3
Yarn micrographs
378
Appendix 5
A5.2.3 Air-jet yarn
The twist triangle in air-jet spinning is short, the tape of entering fibers is wide, and
the emerging yarn is temporarily very hairy. These hairs are then rearranged and
embedded by a second twister to give the structure peculiar to air-jet yarns. The hairs
are wrapped around the core and provide the forces that give the yarn cohesion. Hairs
have to be long enough to give a reasonable probability of the outermost ends becoming
anchored in the structure during the second twisting process so that they act as
binders. With 1.5 inch 1.5 denier (or finer) man-made fibers there is little difficulty
in this respect, but there are problems with short cottons. Longer cottons can be used,
and a photomicrograph is shown in Fig. A5.3(b). The yarn was made from 1.125 inch
of Californian cotton; the wraps can be seen clearly where marked W.
A5.2.4 Rotor yarn
The structure of rotor yarn is controlled mainly by the lying of fibers on a core that
has both real and false twists. Thus, the structure contains a twisted core but the twist
level varies from core center to the outside sheath. The difference in helix angle can
be seen by the shape of fiber cross-sections such as the one shown in Fig. A5.4(a). In
this case, fibers in the central area (circled at B) were almost round, which indicates
that they were almost parallel to the yarn axis. Those on the outside were decidedly
elliptical (indicated at A) because the fibers were cut at an angle. Wrapper fibers are
produced when the yarn intersects the ingoing fiber stream and an example of the
effect can be seen in Fig. A5.4(a). In general, cross-sections vary according to spinning
conditions and the yarn being spun.
The structure needs higher twist levels than those used for comparable ring yarns.
This reduces productivity and makes the yarn harsher. The percentage difference
Wrapper fibers
Migrating fiber
A
Rotor yarn
Rotor yarn
Slice = 0.035″
thick
B
(a)
(b)
Rotor yarn
Ring yarn
(d)
(c)
Fig. A5.4
Cross-sections of various yarns
Advanced topics III: Staple yarn structures
379
between the twist in the core and that of the sheath is a measure of the change in
structure. According to Deussen [3], the difference in twist for cotton yarns ranges
between 0 and 20%, whereas for polyester yarns it ranges between 10% and 45%.
Fibers migrate in rotor yarn, but not so strongly as with ring yarn because of the
lack of a well-defined twist triangle. Figure A5.4(b) shows a sample view along a
piece of rotor yarn in which sections [4] were made at 30 micrometers apart. To make
presentation easier, readings have been plotted as a polar graph and only a portion of
the total is shown. The thickness of the slice of yarn shown was 0.035 inches. The
total fiber traced a spiral path, had seven coils, and migrated between 1.00 and 0.26
of the yarn radius. Its end was hooked. Of course, this is only one fiber among
millions, but it is hoped that it helps to convey the idea of a typical shape. Figures
A5.4(c) and (d) show a comparison of the cross-sections of rotor and ring yarns made
from the same batch of fiber and spun to the same count. The rotor yarn section
shows peripheral fibers at a relatively large diameter but these are part of a loose
wrapper fiber system rather than hairs. Close examination of the packing density of
the fibers shows that the center of the rotor yarn is more tightly packed than the
outside, whereas the ring yarn is relatively uniform in this respect.
Wrappers are created when fibers entering the rotor are laid on the false twisted
yarn at the take-off point near the rotor groove. The yarn passes in the region of the
fiber entry stream once per revolution of the yarn tail. Thus, there are periodic
wrappers along the length of the yarn. Fibers are collected in a roughly triangular
groove and the prism of fibers collected there is subjected to a high twist as it is
removed from the collecting surface. There is a sort of three-dimensional twist triangle
in which there is some fiber migration and a relative movement of the fibers, which
modify the structure when the wrappers become overlaid on the surface [5]. This
refers to the wrapper fibers mentioned earlier. The zone is diffused, it lies inside the
rotor, and is difficult to recognize. Following that, the whole structure is untwisted as
the false twist is removed. The result is a complex structure with a rather roughfeeling surface due to the wrappers. This affects the hairiness, as illustrated in Fig.
A5.3(c). The picture also shows the wrapper fibers and it should be noted that,
although in general, rotor yarn is more bulky than ring yarn, the hairiness is less.
The structure is readily seen by attempting to untwist some rotor yarn. It will be
found that there is never a state when all fibers are parallel in the untwisted yarn, as
will happen with ring yarns. Either the core of the rotor yarn has some twist when the
outer layers are untwisted, or the outside has reversed twist when the core is untwisted.
Furthermore, when one section is untwisted, the neighboring portion may not be, as
shown in Fig. A5.5; the ends of the portion of yarn shown were restrained from
untwisting by wrapper fibers, whilst those in the center of the picture were almost
completely untwisted to form a ribbon. Differences in yarn structure such as these
make measurement difficult; hence, it is usual to rely on the calculated machine
twist. The reversed twist method (i.e. twist–untwist) is sometimes used and it is also
possible to measure the angle of the surface fibers, although this is a tedious process.
Additionally, elongational straining of the yarns produces changes in characteristics.
Strained yarns perform badly in weaving.
A5.2.5 Self-twist (ST) yarn
Referring back to Fig. A5.3(e), two staple yarns which have been self-twisted are
shown. One yarn was made from black fibers and the other from white ones. The
380
Appendix 5
Fig. A5.5
Composite micrograph of untwisted rotor yarn
section shown is near a zone in the white yarn, which originally had zero twist, with
S twist on one side and Z twist on the other. The yarn tries to relieve the torque by
rotating and lessening the twist on either side. As it rotates about its axis, it is likely
to ensnare fibers from the other yarn and wrap them about itself. Black fibers wrapped
about the white yarn can be clearly seen. Also, white fibers were wrapped around the
black yarn but these are difficult to see in the micrograph, it being remembered that
the zero twist zones of the black and white yarns need not coincide.
A5.2.6 Comparing hairiness of various yarns
Many measurements have been made regarding yarn hairiness but the matter is
complicated because any friction acting on the surface of a staple yarn raises hair
from the surface. In ring spinning, the yarn is usually less hairy in the balloon than
it is when it is laid on the surface of the bobbin. This is because the yarn is scraped
over the traveler. Centrifugal force, arising from the rotation of the bobbin, causes
fibers to stand out from the rotating surface. This creates a bed of outstanding hairs
(rather like the surface of a pile carpet), onto which is wound the newly made yarn.
The yarn twists as it is later removed in unwinding, and the surface is again modified.
This is because some of the hairs become entangled with others from different coils
of yarn on the same bobbin. Movement of the yarn over guides in the winder has a
further effect. Thus, it is not surprising that the hairiness is sometimes quite variable
and that research results vary. Some results from Salah [6] (Fig. A5.6), show that
rotor yarns have a larger body than ring yarns. This supports the finding that the
average packing density of rotor yarns is less than that of ring yarns.
With blend yarns the results became more scattered and this is possibly a reflection
on the blending. Other studies have shown that the population of fibers at various
D √Ne is the normalized diameter of the surface bounding the yarn
45
Rotor
Rotor
40
40
45
AV D√Ne
D√Ne
Ring
35
Ring
30
35
30
25
25
100% Polyester
20
20
3.8
4
4.2
TM
Fig. A5.6
4.4
0
25
50
75
100
Blend, polyester/cotton (%)
Variations in normalized yarn ‘diameters’
Advanced topics III: Staple yarn structures
381
stages of processing varies much more than many people expect. The percentage of
polyester in a polyester/cotton blend had some effect but a trend line could only be
established by taking averages over the range of twist multiple. However, it is quite
clear that the fiber packing density, and therefore the hand and cover, vary substantially
from yarn to yarn. Yarns from different machines of various designs differ, but the
trends are similar. It might be noted that, in presenting results, a source of variance
is normalized by multiplying the diameter of the theoretical cylinder by √Ne. There
is still a weak dependence on count even after normalization. This implies that there
is a small error in assuming that yarn diameter is inversely proportional to √Ne.
Comparison of the cross-sections of ring and rotor yarns shows the ring yarn to be
more densely packed. Despite this, the rotor yarn has a looser sheath structure, which
is bound tightly in various places by fiber wrappers.
A5.3
Yarn behavior
A5.3.1 Frictional behavior
Structure affects the frictional behavior of a yarn. It is known that a hairy yarn has a
different coefficient of friction from a less hairy one. Chattopadhyay and Banerjee
[7] showed that a rotor yarn running over a ceramic guide had up to 20% lower
coefficient of friction as compared to a ring yarn. Increased running speeds sometimes
reduced the coefficient of friction. The particular polyester tested showed reductions
in friction, whereas viscose rayon yarns showed an increase. For a given fiber, the
frictional behavior is affected by the finish applied, or, in the case of cotton, the
natural finishes removed. In the case of wool, natural finish is removed in scouring
and the fibers are oiled; the quantity and lubricity of these oils affect the frictional
behavior. Also, experience has indicated that winding and other processes affect the
surface characteristics of the yarn. For these reasons, no more than an approximate
guide can be given to the frictional characteristics of specific yarns.
A5.3.2 Shrinkage in yarns
Untextured filament yarns shrink very little unless the temperature is caused to rise
above the glass transition point for the particular polymer. As discussed in Appendix
A4.8, thermoplastic textured filament yarns can shrink, and the degree of shrinkage
is determined by how well the yarns are relaxed. Shrinkage and yarn bulk are related,
and this fact is evident in the fabrics made from the yarns.
All staple yarns suffer a small percentage twist contraction, and variations in
fabric finishing and laundering can produce shrinkage. Cotton yarns can be treated
chemically to reduce shrinkage of woven fabric in service, and woven fabrics made
from popular polyester/cotton yarns are reasonably stable in this respect. However, in
single-jersey knitted fabrics, the twist liveliness of a yarn affects the fabric structure
significantly. Relaxing the yarn by steam or water treatments prior to knitting helps
to control this problem.
Wool is a special case. It is a scaly, visco-elastic fiber that has many superior
crease recovery properties but is vulnerable to post-spinning shrinkage. Shrinkage
has been a problem for many years because of the ratcheting mechanism of the scaly
surfaces of wool fibers. One method of reducing the shrinkage is to coat the fiber
with a polymer to smooth over the scaly surface. Rosa et al. [8] showed a micrograph
382
Appendix 5
of a treated wool fiber that illustrates the character of a surface smoothed by a
chemical additive. Of course, care has to be taken not to interfere with the other very
desirable properties of wool. Henshaw [9] mentions several authors who have worked
on the problem of shrinkage and he cites high drafting forces as being another cause.
Fiber crimp tends to be most pronounced in the trailing ends of fibers. The contribution
to bulk and hand varies according to the position of the fiber in the yarn. Fiber
migration is extensive.
References
1.
2.
3.
4.
5.
6.
7.
8.
9.
Smith, De Witt W H. Textile Fibers: An Engineering Approach to their Properties and Utilization,
ASTM Proc, 44, 543, 1944.
Barella, A. Yarn Hairiness, Text Prog, 13, 1, 1983.
Deussen, H. Rotor Spinning Technology, Schlafhorst Inc, Charlotte, NC, USA, 1993.
Patel, J R. TX 591 Project Report, NC State Univ, USA, 1970.
Lord, P R. The Structure of Open-end Spun Yarn, Text Res J, 41, 778–84, 1971.
Salah, H A. Bulk and Hairiness of Open-end Spun Yarn, MS Thesis, North Carolina State Univ,
USA, 1972.
Chattopadhyay, R and Banerjee, S. Frictional Behaviour of Ring, Rotor and Friction Spun
Yarn, 1990+.
Rosa, M J, Gómez, L, Coderch, L and Erra, P. A New Process for Exhausting a Permethrinbased Mothproofing Agent on Wool Fibers, J Text Inst, 82, 4, 1991.
Henshaw, D E. Worsted Spinning, Text Prog, 11, 2, 1981.
Appendix 6
Advanced topics IV: Textured yarn structures
A6.1
Yarn hysteresis
A6.1.1 Internal friction in the yarn
The twisting of the filaments under heat can affect both the real and apparent frictional
behaviors of the yarn. If the surface of the filaments is overheated, the fiber finish
might deteriorate by oxidation or some other process and this can change the actual
coefficient of friction between the filaments. Of course, the heating affects the
morphology of the visco-elastic polymer and this affects both the elastic and the
viscous forces acting within the material. External lateral forces acting on a softened
polymer can cause flats to be formed on the filaments as shown in Fig. A6.1.
All the time that lateral forces persist, it is difficult for the individual fibers to
rotate about their own axes. The filaments try to rotate about their own axes when
twist is removed from (or added to) the yarn, but the flats cause an impediment to this
untwisting (or twisting) process.
The next result of these factors is to cause the torque/twist1 characteristic of the
Flattened
filaments
Fig. A6.1
Filament yarn cross-section
1 Torque may be regarded as the torsional analog of extension. Thus, a torque/twist curve has
many similarities to a load/extension curve.
384
Appendix 6
yarn to have a distinct hysteresis loop, as shown in Fig. A6.2. Some of the energy
used in distorting the material in normal use is dissipated in overcoming friction and
is not available to return the structure to its original shape; this makes the hand of the
yarn feel crisper. Furthermore, it affects the development of bulk as is explained in
the next sections. In addition, the flattening of the filaments changes the luster of the
material and may produce a sparkle.
Overheating the fiber causes it to shrink and to change polymeric structure. The
change alters dye affinity and is sometimes associated with polymer discoloration.
Some fibers, such as acrylics, are prone to yellowing if overheated in an atmosphere
containing oxygen. Deterioration in fiber finish is also expected, and there may be
damage to the surface of the yarn. On the other hand, insufficient heating leads to
improper heat setting and poor yarn performance. Thus, careful control of temperature,
and sometimes atmosphere, is needed to ensure high quality and adequate performance
of the yarn. A non-oxidizing gas, or steam, may be used to control the atmosphere.
Within the thermoplastic classification there are a number of yarn manufacturing
methods available of which one is the false twist method. Also, there are a number of
variations within the false twist category. Systems have undergone a great deal of
development over the years.
Torque
A6.1.2 Visco-elastic effects in the yarn
Consider false twist yarn. Yarn becomes heat set when the twisted yarn is at a
temperature above Tg. (The glass transition temperature is the temperature at which
polymer softens.) As the yarn leaving the false twister is cooled and untwisted, the
individual filaments become stressed. If the filaments are separated and then relaxed,
they occupy a greater volume than formerly. They try to go individually into one of
the minimum energy shapes (e.g. Figures A6.4(a) and (b)). However, they will not
completely succeed in doing so because of interfiber friction and viscous effects
within the polymer. The effects of fiber migration and interference between various
helices and snarls tend to magnify the effects of friction. Frictional and viscous
effects cause the torque/twist curve to take the form of a hysteresis loop. The coercive
torques and residual twists vary throughout the process, as shown in Fig. A6.3. In
stage (a) of the process, the filaments are taken from their original stress-free, straight
Coercive torque
Twist
Residual twist
Assuming no filament buckling
Fig. A6.2
Hysteresis in twisting yarns
Advanced topics IV: Textured yarn structures
385
B
Zero torque
Twist
Torque
A
Phase:
Twist
Heat treat
(b)
(a)
New zero torque
O
C
O
G
Torque
Torque
Twist
F
E
D
Phase:
Untwist
Relax
Shape:
(c)
Fig. A6.3
(d)
Hysteresis in the false twist process
condition to a helical, stressed condition (curve AB on the hysteresis loop). The yarn
is then heated above Tg and cooled again to remove the stress so that the filaments
reach a stress-free, helical condition at stage (b). The yarn is next untwisted and the
filaments go from the stress-free helical condition to a stressed but straight condition
(curve CD on the hysteresis loop in stage (c)). In the last stage, the yarn is relaxed
under conditions that allow filament separation and movement towards the new minimum
energy condition (curve EF). The units for the abscissa become coils/inch (or coils/
meter) instead of turns/inch (or turns/meter). Frictional and viscous forces determine
how far up the curve is the final point. At point E, there is a large torque OE acting,
and this tends to produce a snarl rather than a helix, if the yarn is completely relaxed
(Fig. A6.4(b)). If, however, the yarn is heat treated under proper tension and the stress
is removed at some point F, the tendency will be to produce a helical minimum
energy shape rather than a set snarl (Fig. A6.4(a)). It follows that a second heat
treatment under proper conditions in the zone EF produces a bulky yarn. Lack of
such a second heat treatment tends to produce a stretch yarn.
386
Appendix 6
(a)
(b)
Coil separation
Filament diameter = d
(c)
Coil pitch = D + d
Fig. A6.4
A6.2
Coil diameter = D
Filament shapes
Yarn bulk
Yarn bulk depends on the geometry of the helices and how closely the coils pack
together. Consider the case of adjacent coils of similar diameter, packed as shown in
Fig. A6.4(c). Assume that:
1
2
3
4
There are sufficient lateral forces to keep adjacent filaments in contact.
The fibers exist as coils.
All helices are closely packed.
Adjacent helices differ in geometry, with frequent helix reversal points (which
makes it unlikely that one helix will intermesh with another).
At first assume that the enclosing box for a single helix is a unit of unshared volume
and we will refer to this as a unit standard cell. The average height of the cell is
designated as L (i.e. the coil separation) and the pitch is (D + d ). Hence the average
standard unit cell volume is (D + d )2 L. If, however, some coils do not touch within
the cell, or they are intermeshed, a factor K may be introduced to take this into
account.
Advanced topics IV: Textured yarn structures
387
Thus, the average volume of a practical unit cell is:
Volc = K(D + d )2L
[A6.1]
One coil contains a filament of about πD units in length. If the coil height is h, the
volume per coil is approximately K(D + d)2 h
Volume/unit length of filament ≈ K(D + d )2 h/πD
≈ Z′D(1 + d/D)2
[A6.2]
but (1 + d/D) ≈ 1 and volume/unit length of filament ≈ Z′D.
Volume/unit length of filament is related to the specific volume (volume/unit
mass). The factor Z′ is intended to take into account the obliquity of the coil as well
as the value of K. As an approximation:
bulk ∝ (D cos θ)
[A6.3]
This parameter can be related to the final state on the hysteresis loop (G in Fig.
A6.3). The abscissa in Fig. A6.3 might be thought of in terms of (bulk)–1. The final
position depends on the original torque, the tension and temperature during the
second heating phase, and the characteristics of the filaments themselves. If the cells
intermesh, K < 1 and the bulk might be affected drastically. There is not likely to be
a great deal of similarity between adjacent helices and severe collapse is unlikely. The
model is reasonably valid providing the coils do not flip into the snarled state.
A6.3
Fiber migration in textured yarns
A6.3.1 Filament migration
Filament yarns, when false twisted, theoretically have no net twist when they emerge
from the process. However, it is quite possible for there to be alternating twist with
zero mean. Most of the points along the yarn might have some twist in the filaments,
S twist, or Z twist, or some combination thereof. A typical filament shape is shown
Fig. A6.5, wherein the twist reversal points are indicated.
Fiber migration tends to prevent intermeshing of the coils and keep them separated.
Consequently, the parallel coil structure is realistic. This is important because such a
structure entraps large volumes of air, which greatly improves insulation and
compressibility of the yarn. To reiterate, this gives improvements in hand that are
perceived as the warm, soft feel. An interesting collection of micrographs of textured
yarns is given by Lodge [1]; a variety of coiled, zigzag, and buckled filament shapes
produced by the various texturing systems are shown.
Helix reversal points
Fig. A6.5
Fiber migration in textured yarns
388
Appendix 6
A6.3.2 Snagging and pilling
An undesirable facet of some of the yarn structures when assembled into fabric, is the
tendency for them to snag and pill. A snag relates to yarn withdrawn from the surface
of a fabric to make an unsightly fault in the material. Single-jersey knitted fabrics are
particularly susceptible to this problem because yarn can be withdrawn from a course
with relative ease. The very strength of the filaments ensures that, if a protruding end
is caught on an external object, yarn is withdrawn from the fabric instead of breaking
off at the surface. This might cause a collapse of several adjacent loops of yarn, and
any such distortion spoils the surface of the fabric. One solution to this problem is to
modify the structure of the knitted fabric; another is to degrade the strength of the
filament.
Pilling is the formation of many tiny balls of fiber on the surface of the fabric.
Abrasion of the surface tends to texture the protruding fiber ends or loops, causing
them to form into tiny balls. This problem is common to both textured and staple
yarns made from such man-made fibers. Again, an important factor is the strength of
the fiber. If such pills form with a weaker fiber, they break off during laundering, or
wear away, and the fault goes relatively unnoticed. For this reason, many man-made
fibers and filaments used for apparel are deliberately de-rated in strength to combat
the problem. For this sort of end use, the strength of the filaments is more than
adequate.
Reference
1.
Lodge, R M. How Continuous Filament Bulked Yarns will be Made, 4th Shirley Int Seminar,
The Hague, Netherlands 1971.
Appendix 7
Advanced topics V: Blending of staple fibers
A7.1
Introduction
A7.1.1 Introduction to blending
In the present context the errors are assumed to be random. Blending smoothes
random errors but is not an acceptable solution to the problem of periodic errors.
Most periodic errors are man-made and a reliable solution is to eliminate the cause
of the error.
Consider first the everyday case of blending ingredients in a bowl. With sufficient
mixing it is possible to create an almost perfect blend of the materials in the bowl.
Consider next mixing a series of bowls from a continuous supply containing longterm variations in the proportions of the ingredients. Whilst each bowl might be
perfectly mixed, the mix would vary from bowl to bowl. In a system in which the said
ingredients flow through a single mixing chamber, the output would be smoothed to
varying extents and the mass of ingredients in the mixing chamber would control the
variance in blends from sample to sample. Small samples would be well mixed but
large ones would not be. If the sample were smaller than the mass contained in the
chamber, the variance within the sample would be very low but there would be a
variance from sample to sample. If the sample mass were much larger than that of the
mixing chamber, there would be appreciable variance within the sample. The
irregularities in the blend would be smoothed only for a certain length along the
stream of ingredients.
Forwarding this idea to the textile field, one could consider the supply from the
bale storage, the mill itself acting as a mixer. If we make the impractical assumption
that the mill is a perfect mixer and the very practical assumption that there are longterm errors in the supply, we could come to the conclusion that all the yarn produced
during the consumption of a bale laydown would be acceptably smoothed over something
of the order of a billion yards or meters of yarn. There would be some smoothing over
this length, but it might not be acceptable. There certainly might be unacceptable
laydown-to-laydown variation, which the mill would be unable to cure by any internal
blending scheme. Next consider a drawframe, where the supply would be the cans of
390
Appendix 7
sliver in the creel. Assume the mass of fiber in the creel is, say, 1000 lb (≈ 454 kg),
and the amount of yarn produced from that creeling to be of the order of a million
yards (or meters). Then the mass constant would be related to the 1000 lb (≈ 200 kg)
and the limiting long-term error that could be smoothed would be related to the
million yards (or meters). In rotor spinning there is blending inside the rotor, the mass
in that mixing chamber is of the order of, say, 5 mg. The associated mass constant is
a tiny quantity and the limiting long-term error might be only, say, 6 inches (roughly
150 mm). In the vernacular of this chapter, mass constant is proportional to the mass
contained in the mixing chamber under working conditions and the lengths quoted as
proportional to the maximum error wavelengths that can be smoothed by that mixer.
A7.1.2 Problems in defining a blend
If, say, polyester and cotton are blended, it is easy to define the blend because a
micrograph of a cross-section shows the two sorts of fiber as having distinctive
shapes. If two fibers of the same type are blended, the question is no longer simple.
For example, take the case of blending two cottons. Not only is it difficult to discriminate
between the blend components in a micrograph, but often the input materials are
variable.
In the early stages of production, fibers exist in clumps; the average value of any
attribute varies from clump to clump and so does the variance within the clump. The
material is not very homogenous and it is difficult to define. For a good estimate, it
is necessary to measure enough samples to calculate the average and CV of each
attribute. Processing reduces this sort of macroscopic variation by dividing the clumps
into smaller portions and mixing them. However, processing produces its own variations
and it becomes difficult to characterize a blend with absolute accuracy. The attributes
of a fiber do not vary in synchronism and the CVs also vary in unexpected ways.
Some machines have the function of fractionating the fibers and removing some
of the fractions. An obvious example of this is combing, where the fraction removed,
called noil, has a high proportion of short fibers. This obviously changes the blend.
For example, in cotton processing it is desirable to remove the short fibers, which
have a great variability. In removing fibers in the fractionating process, the distributions
of other attributes also change but not necessarily in synchronism. Thus, for example,
changes in fiber fineness are not necessarily related to alterations in the percentage
of short fibers.
A7.1.3 Definition of efficacy of blending
It is known that ‘blending’ can mean different things, according to which is the fiber
property of interest. Each fiber attribute probably exists in a given zone in the fiber
flow line at a percentage that differs from that of the other zones. Furthermore, the
spectrum of percentages changes along the direction of flow. If sufficient samples are
taken at various times from the flow line, the CV of those data provides a good
estimate of the efficacy of the blending process. If the process is perfect, the CV
would be zero, but if the clumps are incompletely separated or the process changes
the order of certain blend components with respect to others, the CV changes. The
CV of a particular blend attribute indicates how well the blending process has worked
for that attribute.
Advanced topics V: Blending of staple fibers
A7.2
391
Bale management
A7.2.1 Warehouse management
Consider a case where a fiber is bought for a period and stored in a warehouse (which
might be in a broker’s facility or elsewhere). If bales were to be used to meet the
production requirements without thought to the remaining stock of fiber, there is a
strong possibility that the best fibers would be skimmed off first. The stock would
then degenerate in quality as time continues. In the case of man-made fibers, the
fiber makers often take care of the problem by making fibers in large ‘merges’. The
attributes of the fibers are made to change very slowly from merge to merge. With
natural fibers the case is different. For one thing, the raw material is inherently
variable within and between seasons. Any skimming could lead to an inadequate
stock at some point through the year (especially if the fiber that is bought is, on
average, only just good enough for the end use). Buying fiber of minimum quality in
the name of an economical purchasing policy can sometimes be a false economy.
From this, it is clear that the laydowns must be based on the technical figures of merit
of the fibers left in the warehouse. Furthermore, the variance within the stock must
be minimized always, otherwise there would be an impairment of ability to furnish
bale laydowns with acceptably low variance at all times
Natural fibers, such as cotton, are seasonal and large quantities of the material
have to be stored for periods in the order of a year. Thus, there are several factors
concerning the blending of such fibers. Factor (a) is to manage the stocks of fiber in
the warehouse so that those laydowns withdrawn on a daily basis are reasonably
consistent. Lack of control in the warehouse leads to ultra long changes; this causes
problems if the old and new fibers become mixed. The time frame of the drift in
properties is measured in months. Factor (b) is the consistency of fiber quality within
a bale. This is determined by gin practice for cotton, sorting practice for wool, and
industrial practice for man-made staple fiber. The time frame is in days. Factor (c) is
to minimize the CV of the various fiber parameters in a laydown so that periodic
removal of fiber from the bales in turn does not produce error. The time frame here
is in minutes. Even the shortest time mentioned covers a period during which, perhaps,
100 000 yd of yarn are produced – a time that would still fall in the category of longterm error as far as yarn is concerned.
It is common to divide the store of bales into attribute categories to make the
management problem more tractable, because the range of values of the various fiber
attributes is very large. Each bale is sampled, the samples are tested, and the bale is
assigned to one of the attribute categories by using some formula that the mill
operator thinks best serves his or her business. The number of categories depends on
the complexity of the business. Bales of fiber do not have equal value, and storage of
poor quality or high cost bales with little chance of their being used is a financial
burden. It is desirable that the rate of movement of bales of each category be logged
so that slow moving categories can be eliminated, unless there is special reason for
having them. Management of the fiber warehouse requires a knowledge of the technical
figures of merit and their mean values over long periods. It also requires the use of
financial figures of merit that take into account the fiber and storage costs, as well as
the value in yarn form.
Warehouse management becomes a problem of maintaining constant proportions
of the fibers that fall in carefully defined categories. Typically, each category contains
only fibers of a certain range in fiber fineness, length, and color attributes. Other
392
Appendix 7
fiber attributes may be added but the number of categories expands exponentially
with the number of attributes controlled. A large number of categories make commercial
application very difficult. Hence, the bales issued for a given laydown can only be
defined within limits and there is bound to be some variation within bales drawn
from a given category. Blend proportions can be defined for any single fiber
characteristic, but they do not all behave in the same way. For example, if one is
interested in fiber strength above all else, then the blend must be arranged to minimize
CV of fiber strength. If, as is more likely, the need is to reduce barré, then the blend
should concentrate on the reduction of the CV of factors that affect color. These
factors include not only the normal color measurements such as yellowness (+b) and
reflectance (Rd), but also factors that affect dye affinity such as micronaire. A further
factor is the surface structure, which can affect the perception of color.
A7.2.2 Bale laydown
The quality of a blend is determined by (a) the fiber buying policy (cost and fiber
quality), (b) the testing and application of the incoming raw material, and (c) the
treatment of reworkable waste being returned from downstream processing. It should
be pointed out that the affordable cost depends on what the end product will bear. The
bales selected for the laydown should depend on the application. As mentioned, the
pattern of laydown is important to get not only good performance of the mill but to
maintain a consistent inventory of fiber in the warehouse. Apart from variations
inherent in normal production, there are particular changes that have to be managed
with care. The latter are referred to as merge changes and they occur at crop change
time in natural fiber production. Merge changes from the production of fibers in a
synthetic fiber maker’s plant occur when significant alterations are made to the
process or product. Every care has to be taken to minimize the possibility of a
customer mixing lots of yarns containing dissimilar fibers. The result of uncontrolled
fiber application and flow will be barré in the final fabric. This is especially true
when significant step changes in important fiber properties are involved. Many
complaints from customers relating to finished fabrics have their origins in yarn
production. Settlement of claims often involves the yarn maker paying for fabric
production and finishing of defective material if the yarn is faulty. An expensive
error!
To help in reducing the variability of the blend components, it is now possible to
use (HVI) and AFIS testing to measure fiber properties on a mass production basis.
Discussions on AFIS and HVI are given in Section A4.7.3, and also further discussion
on HVI appears in Section 11.2.1. Software programs to manage the laydowns
appropriately can further augment results from such testing. A leading example of
this is the EFS software produced by Cotton Incorporated. Table A7.1 gives examples
that compare the values in bale material to those measured in sliver from the same
laydown slice. The comparison shows that the CV, and therefore the blend evenness,
was not always better after carding. In this particular case, sliver samples were taken
at 10 yard intervals (equivalent to perhaps 1000 or 2000 yards of yarn) but these
cannot be regarded as ‘long term’ relative to the bale samples. Consequently, the data
should be interpreted to mean that the relatively short-term variation in many of the
blend attributes, usually deteriorates in carding. The idea of a single ‘blend efficiency’
to describe performance is seen to be misleading because one must concentrate on
the blending of components that matter for the given product. Also, the short- and
Advanced topics V: Blending of staple fibers
Table A7.1
Bale*
Sliver
393
Percentage coefficients of variation
MIC
UHM
STR
ELO
Rd
+b
SFC%
3.3
2.4
0.7
1.5
3.1
4.3
4.7
5.4
4.6
8.0
4.8
6.7
11.1
17.3
Note: * = between bales. The acronyms are defined in Section 5.8.2.
long-term variations have to be balanced to give minimum trouble in year-long
processing.
Of course, there is a wide range of variability in natural fibers according to the
type of fiber, growing area, climate, and seasonal changes, and some bales are more
variable than others. The best summary of the position is that variation within the
bales is not negligible. Here, it is only possible to cite a given case and no representation
is made that it is typical; the purpose is to demonstrate that the problem is significant.
The case in point resulted from a study in which the bales were specially selected to
give as low a CV in micronaire as possible. The choice was made using test results
from many bale samples. Space precludes giving full details but suffice it to consider
a sample of nine bales, as shown in Table A7.1. Variances were averaged for each
horizontal slice of all the bales in the laydown (the reader is reminded that variance
is proportional to the square of CV). The sliver figures are based on the average
variances in samples of card sliver taken systematically over the period over which
the bale slices were consumed. The sliver data include the variances between bales,
within bales, and any effects caused by processing. Upper half mean length varied
little in this case. However, short-fiber content was high and there was a substantial
within-bale variance, or processing had produced an extra variation, or both apply.
This is important because short fibers cause instabilities in roller drafting, which add
to the CVs generated in later processes and thus degrade the final product. The CV
of micronaire was variable despite attempts to control it by selecting bales on the
basis of the cotton broker’s data. Micronaire is important not only because of varying
cross-sectional size, but because of varying wall thickness in immature fibers that
sometimes occupy the low micronaire portion of the distribution. It is possible that
carding exercised a fractionation function and removed some fibers of high or low
micronaire values and thus reduced the CV in that attribute. High CV of micronaire
has come to be recognized as one of the causes of barré. The substantial difference
in the reflectance of the fibers (Rd) is difficult to explain by processing. Perhaps the
values are linked to other attributes vulnerable to change by mechanical processing
or perhaps some fiber crimp is removed and this affects the reflective capabilities of
the fiber. Again, this is only anecdotal and the values quoted should not be taken as
typical for all cases. The points are that the within-bale values are not negligible
compared to the between-bale figures and processing can affect the results. On a
number of occasions the author has observed variations of the different fiber within
a bale that are of the same order of magnitude as those between the bales.
Another set of circumstances sometimes confronts a yarn supplier to the knitting
industry. The product is often judged in the greige state and fiber color becomes
important. In such cases, variability in the +b and Rd values assume greater importance.
(Greige refers to fabrics in the state that they leave the loom or knitting machine [1],
and by extension, it refers to the yarn used in making such fabrics.) In the case of
acrylic yarns, yellowing of the fiber might be a factor.
394
Appendix 7
Obviously, a prime requirement is to assemble a laydown with as little variance
between bales as possible. Since the variability of one fiber attribute does not necessarily
match that of the others, it is necessary to set up a priority system. Each fiber
attribute is given weighting calculated on the end use of the yarn, so that there results
a figure of merit customized for the particular product. To be workable, the figure of
merit must have many combinations of fiber attributes that satisfy the requirements
for a given end use.
A7.3
Mixing in the blow room
A7.3.1 Mixing the fibers flowing through the opening line
In a bale laydown, only the fibers from a limited number of contiguous bales can be
mixed before the fibers pass to the cards. Assume that there is a moving zone that
includes a portion of one or more rows containing (m = m1 + . . . + m2) bales shown
shaded in the simplified bale laydown in Fig. A7.1. Also, assume that only the fibers
removed from the bales in this moving zone are intermixed. Bales from outside the
zone are assumed not to participate in the mixing until the moving zone encompasses
them. Participation stops when the moving zone passes by. The bale plucker moves
slowly down the line of bales in the direction shown and the zone trails the cutting
head. When the cutting head reaches bales 1/2 or Y/Z,1 it reverses. It slowly reciprocates
between bales until the demand temporarily ceases, or until the fiber flow system
calls a halt, or until the laydown is consumed. The moving zone always consists of the
horizontal slices taken from the last m bales passed by the bale plucker as it moves
to and fro. Some systems take slices at an angle, but the idea expressed is still similar.
The fiber passes into a series of mixing zones that consist of several elements. Each
machine in the opening line causes a degree of mixing between adjacent volumes of
moving fiber and even turbulence in the ductwork contributes to the process.
Where laydowns are formed with several bales set side by side to form a pattern
w bales wide and m rows long, averages are taken for each row to yield m average
rows. Fiber from these rows progresses through the system in line astern along the m
direction. In theory, it is assumed that a step change exists in fiber characteristics as
the cutter of the bale plucker leaves one row of bales and passes to the next. It is
further assumed that the fibers from all the bales in a row are adequately mixed. The
plucking head moves on a fairly regular, periodic basis, but there are occasional
1 3 5 .....
2 4 6 .....
Bale plucker head movement
m1 .....
UWY
..... m2
Bale laydown
Fig. A7.1
VXZ
Bale laydown
1 Z often is greater than 40, and the number tends to grow as the technology improves.
Advanced topics V: Blending of staple fibers
395
dwells to keep the fiber flow in synchronism with the demand by the cards. However,
the flow may be considered to be more or less continuous and the characteristics of
the fiber delivered can be considered to vary periodically.
A7.3.2 Errors due to fiber removal from the bales
The characteristics of the fibers can vary cyclically along the fiber flow path because
of the reciprocation of a bale plucker over a bale laydown with varying bale to bale
fiber characteristics. The effect is worsened by any inadequacy of a mixing machine(s)
in the opening line. A peak in the spectrum occurs typically at an error wavelength
related to one traverse of the bale plucker. It is useful to express this error in terms
of a length of card sliver delivered from one of the cards connected to that opening
line. Here, this length is defined as the amount delivered during the time it takes for
the plucker to complete one cycle of its travel. This value is of the order of 1000 yards
with current machinery. The result may be represented in the frequency domain in the
manner of a spectrogram. Some examples of practical results are given later.
A further problem exists. Bales are not uniform throughout. The profile of fiber
characteristics from one cut across the laydown is not the same as another taken at a
different time. The profile varies continuously as the laydown is consumed. This
means that there can be some extraordinarily long errors from the system. Even if the
order of presentation to the spinning machine is scrambled, unlike yarns will be
produced on adjacent spindles. This is a recipe for barré.
A7.4
Theory of blending capacity
A7.4.1 Dispersion in the flow through mixers
The foregoing discussion has indicated that variations can occur over a range of
processing times varying from a few minutes to a year. The opening and blending line
can only deal with changes that occur at less than a certain characteristic time peculiar
to a given installation. As an analogy, consider water flowing down a river into a lake,
which discharges into another river and out to sea. Under steady conditions, certain
levels become established between the two rivers and the lake. If a sudden deluge
causes a rise in the river entering the lake, the effect is not passed on in an unchanged
way to the downstream (discharge) river. This is because the volume of the lake
absorbs some of the sudden rise. The discharge river rises much more slowly. A
similar situation occurs in the opening line; a sudden change in one of the fiber
parameters in the bale laydown is not immediately passed on to the sliver emerging
from the card. The intervening volume and the degree of variability in the fiber flow
control the output.
Consider the flow of fiber passing into a reservoir, as shown in Fig. A7.2(c).
Assume an incremental volume of new fibers (which we will call fiber1) enters the
mixing volume Q and is immediately mixed with all the other fibers (i.e. fiber2) in
that main volume. The excess is ejected and contains the same proportions of each
fiber as exists in the fixed volume Q.
Let the total volume of fiber1 derived from a single pass over one bale be Qin; the
rate of bale plucker movement be V bales/unit time, and the fixed volume be Q. When
the bale plucker passes from one bale to the next, a front is created by the step change
396
Appendix 7
m bales
1 bale
Fiber1
(a) Input
m bales
1 bale
Fiber1
(b) Output
Volume Q
Input pulse
Output pulse
(c)
% Mixing
100
80
60
Q = (1 × Q in)
40
Q = (2 × Q in)
20
Q = (4 × Q in)
0
0
1
2
Volume delivered to mixer (units of S)
(d)
Fig. A7.2
Volume delivered to mixer in units of S (S = normalized fiber volume supplied
to the mixer).
in fiber characteristics. The flow of fiber between successive fronts is called a fiber
pulse. If the normalized flow of fiber is constant at S, then it can be shown that the
volume of fiber1 in the mixer increases exponentially as the new front passes through
it. Normalized fiber flow = S = V Qin /Q. The percentage value can be expressed by
the equation:
Qm1 = Q(1 – e–s)100%
[A7.1]
where Qm1 is percentage of fiber1 in the mixer.
The greater the volume Q, the smaller is S, the slower is the rise in percentage of
fiber1 in the output and the longer it takes to approach the limiting value of 100%.
Thus, a sudden change of fiber, such as is met when the offtake from the bale
laydown passes from one bale to the next, causes the mix to change. The output pulse
leaving the mixing zone is modified by the mixing process and is diffused.
In one of the theoretical examples plotted in Fig. A7.2(d), the mixing volume is
four times the proportion of the bale slice removed as the offtake mechanism passes
over a single bale. It would barely reach 50% of fiber1 before the offtake moved past
the bale. Equation [A7.1] is applicable only while fiber1 is being supplied. When the
Advanced topics V: Blending of staple fibers
397
supply of fiber1 is replaced by that from the next bale in the laydown, the percentage
of the fiber that had originated from the bale just passed declines exponentially, and
the contribution from the new bale begins to rise. Two diagrams illustrate the idea.
Figure A7.2(a) shows a rectangular input pulse, which represents the bale slice being
removed. Figure A7.2(b) represents the percentage of fiber from that slice appearing
in the output. Mixing blurs the boundaries and elongates the volume that contains
some component of fiber1. The rate of change and the length over which the fiber is
distributed depends on the size of the mixing volume and the efficiency of mixing.
For our purposes, the larger the volume Q, the better, because it blurs the boundaries
between consecutive zones in the flow line and thus improves the local blending. The
length over which distribution occurs may be thought of as a sort of mass constant.
In practice, all machines in the opening line contribute a mixing volume, each of
which works similarly. Each machine contributes its quota to the mass constant of the
whole line.
A practical opening line would have no difficulty in blending a single bale slice
with its neighbors. The point of the exercise was to show how fibers from a subject
bale are spread amongst its neighbors in a mixing operation. More to the point is how
far fibers from a subject bale can be spread. A modern bale plucking machine travels
over a laydown causing a cyclic removal of fiber. The corresponding period in the
sliver is usually related to twice the number of bale rows in the laydown. To eliminate
periodic variations in fiber attributes, it is necessary to mix the fiber from a bale
being worked with all the others. This implies that the mass constant of the opening
line should equal or exceed the amount of fiber removed in a single pass over the
laydown. For example, if the mass constant of an opening line is 100 lb and the bale
plucker removes 2 lb per bale slice, the size of the laydown should be no larger than
50 bales. Many operators try to achieve synchronism of making a new laydown
correspond with the work shift schedule by increasing the number of bales in a
laydown. The result of this might be the introduction of a periodic variation in fiber
characteristics with a period equal to one round trip of the bale plucker. Also, if the
bale slices are heavier, a similar situation will arise. Either circumstance is to be
avoided if possible. It has to be recognized that there are limitations to the possible
courses of action. For example, the mass of an average bale slice might be (say)
2.5 lb (assuming a throughput of 800 lb/hr and a maximum rate of bale plucker
movement of over 320 bales/hr). Continuing the example, let there be 100 bales in the
laydown. If we use the criteria just suggested, the mass constant (blending capacity)
would have to be increased to 250 lb.
The discussions above have assumed a steady delivery from the bale pluckers.
However, most bale pluckers stand idle at times, waiting for demand to catch up with
output. Anything that can be done to keep the bale plucker in full use, or to decrease
cutting depth, helps blending efficacy.
A7.4.2 Dispersion in drawing and combing
The fiber flow through the mill might be considered analogous to flow through a
pipeline. The system starts with a reservoir containing, perhaps, 25 000 lb of fiber,
and flows through an opening line at roughly 1000 lb/hr. This assumes that 50 bales
were used in the laydown. The flow branches to supply the cards and each card works
at roughly 100 lb/hr. The card sliver is gathered in cans and forms a secondary
reservoir behind each drawframe. The secondary reservoirs might each contain up to
398
Appendix 7
500 lb. Before combing, there is yet another reservoir, which we will call the tertiary
reservoir, which contains, say, 1500 lb.
Ingoing material is doubled in drawing and lapping which reduces the variance.
Consequently, according to some practitioners, there is a reduced need for good
blending at the early stages. With all the doubling that occurs up to finished sliver,
fewer than four or five bales/creel are involved. Whilst it is true that the doubling and
mixing at these stages helps within that compass, it certainly cannot be completely
effective, especially when viewed against the 25 000 lb or so in the bale laydown. The
idea of mass constant may be applied similarly to that applied to the blowroom. Also,
it should be noted that the ratio of creel mass in the drawframe to the laydown mass
is only of the order of 5:100. Clearly the value is so low that the creeling can only
have a very small impact on extra-long-term errors. Even with multiple-sliver processes,
the effect of step changes in bale laydowns still cannot be offset and this explains why
consistently good blending is needed in the early stages.
A7.4.3 Channeling
Under certain circumstances, differences in fiber supplied to one card compared to
that supplied to another can cause barré problems, and so can differences in performance
of one card compared to another. The problem arises when a drawframe and the
following processes are fed exclusively from a dedicated set of cards. If the mean of
the product from one set of cards differs from that of the others, the yarns differ, and
when they are mixed in winding there can be undesirable step changes in yarn
properties. This phenomenon is known as channeling. A solution is to cross-feed the
drawframes and combers in a pattern likely to minimize the variations. Often color
banding of the cans is used to determine their destination.
A7.4.4 Periodic variations
Assuming that relative short wavelengths of periodic error are controlled by good
inspection and maintenance, there still remains the question of ultra long-term errors.
Difference from top to bottom of a laydown and differences between laydowns produce
a mixture of random and periodic errors. The doubling is effective in reducing the
errors within the limits already described, but doubling is ineffective in cases where
all cards within a set produce a periodic error of the same frequency. Ultra long-term
errors arising from the bale laydown could affect all slivers produced by that opening
line.
Doubling of periodic errors in the creels of subsequent machines produces a
vector addition of errors of similar frequency and amplitude but of random phase.
Phase is determined by the relative longitudinal positions at which the sliver ends are
laid in the creel. Such additions of periodic errors can result in output error components
varying from zero to m times the periodic error originally in the sliver, where m is the
number of doublings. The problem is that these errors are so long that they are rarely
detected. The probability of such an error at a significant level is of the order of one
in several hundred creelings. This means the problem shows itself rarely, but when it
does, the error can be significant for the duration of that one creeling.
When there are periodic variations of different frequencies in the slivers being
creeled, the waveform of variation is complex. The various frequencies beat together
and produce a difference frequency, which may be expressed as a very long wavelength.
Advanced topics V: Blending of staple fibers
399
Thus, there are infrequent periods when the vectors are aligned. At these times, the
amplitudes of the component errors add together algebraically on a temporarily regular
basis. Errors at these times are large and similar in amplitude to the case just mentioned.
Doubling does not diminish periodic errors reliably; thus it is prudent to avoid this
sort of error. Hence, sources of periodic error should be diagnosed and the fault
eliminated as far as possible.
A7.5
Fiber migration and blending
A7.5.1 Longitudinal fiber migration
If the fiber flow consists of large fiber clumps, the blend will be uneven because of
the variations between them. At the beginning of the process, there are large migrations
and very considerable blend irregularities. Unevenness in mass distribution is created
when the blend is uneven and there is fiber migration. The effects of this are offset
by the doubling that occurs at each condensation stage. Consequently, quite large
errors might be generated, even if they are later reduced by doubling. Doubling will
not reverse the migrations and the composition of fibers flowing through the system
will show effects of both types of process.
The equation of flow can be considered as a spectrum of sinusoidal variations. Let
one of these sinusoids have a wavelength of λ, and let the fiber migration be m, where
both variables are measured in consistent units of length. Figure A7.3 shows a simple
Short fiber
Remaining fiber
(a)
Constant linear density
Short fiber
(b)
Varying linear density
Remaining fiber
Short fiber
(c)
(d)
Error amplitude
Fig. A7.3
Effect of longitudinal fiber migration
Linear density
Remaining fiber
400
Appendix 7
sinusoidal blend error in a perfectly leveled strand. In Diagram (a), the leveled strand
has a constant overall linear density, but the two fiber components vary sinusoidally
in a complementary fashion. The two components are shown as short fibers and the
remaining fibers respectively. The short fiber has a linear density of n1 + A sin α,
where A is the amplitude of variation in linear density of the component and α is the
length along the flow line expressed as an angle. For explanation purposes, let the
curves be transposed as in Fig. A7.3(b). To account for the longitudinal migration, the
remaining fiber portion is moved horizontally to the right as in Fig. A7.3(c). Diagram
(d) shows the addition of the ordinates for short and remaining fibers, and this
represents the overall linear density after migration. Clearly migration has changed
the overall linear density so that it is no longer level. The amount of error introduced
by this mechanism depends on the changes in the blend ratio; if the blend had been
perfectly even, no amount of migration would have produced an effect. In practice,
there are many wavelengths of error and the amount of error produced varies; more
complex analysis is required but the exercise has demonstrated how the interaction
occurs.
Expressing the results mathematically, the phase change, φ, due to the migration is:
φ = 2π (m/λ) radians
[A7.2]
The fiber moves (φ/2π)λ length units. The top curve in Fig. A7.3(c) is represented by:
n1= y1 + A sin α
[A7.3]
and the bottom one by:
n2 = – y2 + A sin (α + φ)
[A7.4]
The curve in Fig. A7.3(d) is the difference between Equations [A7.3] and [A7.4].
nm = k + 2A(cos(α + φ/2) sin φ/2)
[A7.5]
where n1 is the linear density of the top portion, n2 is the linear density of the bottom
portion, and nm is the linear density of the strand after migration. The values y1 and
y2 represent the mean values of a long length of the two portions and k = n2 – n1.
This means that the error introduced in linear density is a function of the error
wavelength of the interface and the phase change referred to earlier. One of a number
of critical phase changes is when the fiber migration equals half the error wavelength
of the interface. The effect is worse when sin φ/2 = 1.0 (for example, when the phase
change is 180°). However, there is no effect on linear density when the phase change
is zero or any multiple of 2π radians (360°).
To repeat, the practical importance of this is that (a) there has to be significant
fiber migration, and (b) there have to be differences in the blend proportions along
the strand, before errors in linear density are caused by the phenomenon. Generally,
the greater the gradient of the change in a blend component along the length of the
strand, the greater is the change in linear density of the strand. Therefore, sudden
changes in blend should not be allowed to happen; changes should occur very slowly
if problems are to be avoided. As a practical matter, this means: (a) keeping the fiber
clump sizes as small as possible at every stage, (b) arranging to avoid dissimilar bales
being in proximity in the bale laydown, and (c) promoting as much mixing as possible
in the largest mixing volumes possible. There is also a requirement that the average
value of the moving zone shown in Fig. A7.1 should vary by a minimum amount and
this requirement may conflict with others.
Advanced topics V: Blending of staple fibers
401
A7.5.2 Incremental effects of migration on staple fiber blends
For simplicity, consider a two-component system, shown in Fig. A7.4, in which
Component B contains fibers that all have a common fiber attribute and Component
A contains the rest of the fibers. However, let the proportions of the two components
vary along the length. A migration of one component with respect to the other along
the line of flow causes blend changes with respect to all fiber attributes. For instance,
if fiber length is a major variable, then the fiber migration will change the distribution
of fiber lengths along the strand.
In Fig. A7.4(a), the two components are labeled A and B. The masses of fiber at
any position x from the zero point in the central zone are described by two equations.
This zero point is shown by a black dot on the o–o line. The ordinate reference line
is o–o and the total mass at position x is ( y + y1). The local blend proportions are b
= y/(y + y1) of Component B and b1 = y1/(y + y1) of Component A. In Fig. A7.4(b),
Component B has migrated to the left by an amount φ and we shall consider only the
zone Z. The equation for y1 is unchanged but the shift of origin causes the other
equation to become:
y = m(x + mφ) + b
The blend proportion for component B after migration becomes:
bm = (y + mφ)/(y + y1 + mφ)
and when mφ is small with respect to ( y + y1)
bm ≈ b + mφ/(y + y1)
[A7.6]2
A
y1 = m1x + b1
b1
O
y = mx + b
b
O
B
(a)
Direction x
Direction y
A
Z
O
O
B
φ
(b)
Fig. A7.4
2 Differentiating, db/dφ → m(y + y1).
y = m(x + φ) + b
Fiber migration gradient
402
Appendix 7
Equation [A7.6] may be interpreted as implying that the rate of change of the blend
proportion due to migration depends on the gradient of the components.
In other words, if the entering blend component is of variable mass along its
entering length, there will be changes in blend proportions in the output due to the
migrations caused by drafting. This is distinct from the linear density of the total
strand. As before, if the entering blend components are level, no amount of migration
will cause a blend change.
A7.5.3 Blending by migration
If the blend proportions vary along the length of the strand, the migration causes
changes. Thus, the drafting of an irregularly blended but perfectly even strand not
only causes irregularity in output linear density but also causes changes in blend
proportions. Fibers leaving a machine that drafts the fiber are in a different order
from those arriving. Sequential cross-sections contain samplings along the flow path
of the input and output material and the cross-sections vary in content from one to
another. Calculating the linear densities of the two components from Equation [A7.3]
and taking the ratio of the two components, gives us the blend proportions. The
average original blend proportion is taken as 0.5 in this example and φ is the phase
angle representing the longitudinal fiber migration. The effects are cyclic and at
certain phase angles the variation in blend becomes a minimum, as shown in Fig.
A7.5. An actual blend contains a spectrum of cyclic components and, for a given fiber
migration, there is a spectrum of relative values of φ. The result of this is that some
cyclic components are emphasized in comparison to others and there is a sort of
resonance pattern. Thus, while the overall effects of migration tend to improve the
blend, strong patterns of variability can arise which can have deleterious effects in the
subsequent processes and products.
A7.6
Real blend variation
A7.6.1 Variations in mechanical attributes of fibers
An illustration of the levels of variability of mechanical attributes was given by a set
of experiments carried out at Cotton Incorporated. About 7000 yards of sliver were
measured at 10 yard intervals along that length for the various fiber parameters. For
the examples quoted here, it was preferred to use periodograms (similar to spectrograms)
Relative blending ratio
1.0
φ = 0°
0.5
0
φ = 30°
0
Fig. A7.5
360
α (degrees)
720
Phasing of blend components
Advanced topics V: Blending of staple fibers
403
0
leng
th,
200
100
le (
yds
)
(a) Short-fiber content
log
sca
0
400
Wav
e
Leng
200
100
2000
4000
th fr
o
slive m the s
r sa
mple tart of th
e
(yds
)
FFT amplitude
rather than frequency based curves. Each parameter yielded a time series to which a
fast Fourier transform (FFT) was applied to change it into the frequency domain (or
wavelength domain). To get the range of wavelengths to describe the phenomena, it
is necessary to use a log scale in a fashion similar to that used with spectrograms. To
lessen the confusion from the multitude of results, only short-fiber content and
micronaire are quoted here. Variations in the short-fiber content are shown in three
dimensions in Fig. A7.6(a). The calculated value for error wavelength of the bale
plucker cycle was about 700 yards; that compared well with one of the major peaks
in the practical values. There were quasi-random variations denoted by the ‘mountains’
in the diagram. These illustrate the danger of applying an FFT to a limited time
series; often-anomalous results can be obtained. However, there were also periods
when a strong FFT component was noted at the wavelength corresponding to the bale
plucker movement. The strength of these peaks varied along the length of the material
and this, in part, might be due to the intermittent demand on the supply to the mixer.
FFT amplitude
2000
Card sliver
Drawn sliver
1000
0
100
1000
Error wavelength, log scale (yds)
(b) Micronaire
Notes: Input micronaire is the average of the values from 8 creel slivers.
Error wavelength for drawn sliver expressed in terms of card sliver.
Fig. A7.6
(a) Three-dimensional view of short-fiber content; (b) Comparison of micronaire at the
input and output of a drawframe
404
Appendix 7
The length of sliver depicted took about an hour to produce, during which time about
5% of the bale laydown was consumed.
Micronaire values affect the dye performance of the yarn. A typical pattern measured
with card sliver is given in Fig. A7.6(b). As expected, a distinct peak was present at
the bale plucker cycle frequency. These variations showed as color barré in fabrics
made from the yarn produced from the sliver mentioned. The barré had a periodicity,
which related exactly to the cycle frequency just mentioned. The FFT analysis was
meaningful and accurate in this case.
A blend can be judged in a number of ways. Consider the real case of the 100%
cotton sliver shown in Table A7.2, which has been selected because it was a rare,
almost worst case scenario. The scenario arose from a badly organized commercial
bale laydown and it was compounded by the lack of a blending machine in the line
within the particular industrial plant. Card sliver was sampled every 10 yards and the
samples were tested for various fiber attributes.
The drawframe was creeled with 10 card slivers in the input and the draft ratio was
10. It may be recalled that, according to doubling theory, the ratio should have been
√10 in favor of the output. As might be expected, the actual output sliver had considerably
larger values of CV than the theoretical values (Table A7.2). The output did not vary
in sympathy with the input, despite the synchronization of the testing. Similar experiences
have been had in carding and opening. Correlation between input and output has been
determined to be statistically insignificant. This is because of the differential longitudinal
migration of fibers in the fiber stream within the process concerned.
A7.6.2 Coefficients of variation
It is possible to use coefficients of variation (CV) or variance in a specific fiber
characteristic, for example, if one is trying to control fiber micronaire or short-fiber
content (because of drafting problems). However, each of the spectra for different
fiber attributes exhibits a different characteristic. Figure A7.6(b) compares some
spectra before and after the first passage of drawing.
The draft in drawing was moderate, and the degree of fiber migration small. The
resulting effects were also small and thus drawing was found to produce only a
modest effect on wavelength distribution of each of the many fiber parameters actually
tested. Opening and carding produced irregular peaks, which sometimes corresponded
to the bale plucker movement.
A7.6.3 Color variations
Standard colorimetry can be used to measure color of loose fiber, sliver, roving, yarn,
and fabric. In modern mill practice, fiber color is measured on HVI equipment and
it is expressed as yellowness (+b) and reflectance (Rd). In staple yarn mills, it is rare
Table A7.2
Variations in sliver (%CV)
Card sliver (drawframe input)
Theoretical output value
Drawn sliver (drawframe output)
MIC
UHM
UI
STR
ELO
SFC
3.0
1.0
3.1
1.7
0.5
1.8
1.1
0.3
1.1
3.4
1.1
3.6
5.2
1.6
4.2
11.4
3.6
11.9
Note: The acronyms are defined in Section 5.8.2.
Advanced topics V: Blending of staple fibers
405
to measure the color of yarn or the intermediates in this respect, although this might
change because of the growing availability and use of optical measuring devices. In
the following text, some preliminary data relating to measurements made on yarn and
intermediate products is given in the hope of providing some guidance on the
possibilities.
A7.6.4 Color measurements on greige fabric
Some knitted fabrics were made from the carded sliver referred to in Fig. A7.6(b).
Sliver-to-yarn spinning was used to avoid any distortions from the processes of
drawing and roving. Also, the samples of sliver actually knitted were contiguous to
the samples used for HVI testing. A sample of the results at a color wavelength of
700 nm (i.e. red) is given in Fig. A7.7. Similar curves were produced for other color
wavelengths. The error wavelength were determined by applying a Fast Fourier Transform
(FFT) to the time series of measured results.
The similarity of the spectra implies that micronaire is related to the color response
in fabric form. It is known that the dye uptake of cotton fibers is affected by the
micronaire values and therefore it might be expected that the changes noted would
show up even more clearly in dyed fabric.
The exaggerated data shown here imply that bale-to-bale variation of fibers in the
bale laydown affected the color of the product cyclically and that blending in the blow
room had been inadequate. This latter fact was true for the particular case but it must
be emphasized that it was a departure from normal, modern practice.
A7.6.5 Color measurements on sliver
Further blending reduces variations in blends but complete homogenization is not
possible. Thus, it should be expected that some variations occur in all the fiber
attributes when measured in yarn or in any of the intermediate products. This observation
FFT amplitude (arbitrary scale)
1500
1000
Carded sliver
500
0
100
Fig. A7.7
Greige knitted fabric
Error wavelength, log scale (yds)
1000
Comparison of color is card sliver and fabric made therefrom
406
Appendix 7
Cyan
Mean = 4.2%
St dev = 1.2%
Yellow
Mean = 12%
St dev = 3.3%
Y
C
Magenta
Mean = 2.2%
St dev = 0.7%
M
Gray
Mean = 3.9%
St dev = 1.0%
K
200
250
Energy absorption (arbitary scale)
Fig. A7.8
Comparison of color components in sliver
applies to the color spectra as well as to the other attributes. It is possible to measure
the color spectra fairly easily and this provides a useful view on the efficacy of
blending.
The color spectrum of a fiber assembly may be expressed as color separations,
such as those used in the printing of color photographs. One method is to use the
cyan, yellow, magenta, gray (CYMK) system of color values and to quote the color
depths of the textile material, such as that of some sliver as illustrated in Fig. A7.8,
instead of the Hunter scale of +b.
Examination of scanned images of various sliver and yarn specimens indicates
that there are micro-variations in yellowness, which are likely to arise from incomplete
homogenization of the fibers within the blend. Even combed sliver shows striations
in the yellow separation, which are not very visible in the other color components.
More work is needed to comprehend the longer-term ramifications of such anomalies.
Reference
1.
Beech, S R (Ed). Textile Terms and Definitions, The Textile Institute, Manchester, UK, p 114.
Appendix 8
Advanced topics VI: Drafting and doubling
A8.1
Theories of drafting
A8.1.1 Purposes of drafting
The purposes of drafting are to elongate the strand (a) to change the linear density,
and (b) to improve the fiber orientation within the strand. Two types of drafting are
common, namely roller and toothed drafting. Roller drafting will be dealt with first,
but many of the remarks apply to toothed drafting as well. The combing roll used in
rotor spinning typifies toothed drafting.
Roller drafting is known to create irregular fiber flows, particularly when there is
a large variation in effective fiber length. Thus, errors are introduced into the output
strand by the act of drafting. The means of control for this irregular flow include the
use of aprons, pressure bars, and rolls. All of these control elements attempt to
restrain the forward movement of the fibers within the draft zone until the last
feasible moment. More detailed discussion of this is given in Chapter 3. The ideal
control element acts to keep the floating fibers at the speed of the back rolls. It only
permits the fibers to accelerate to the front roll velocity when the leading end nears
the front roll nip. This means that the control element is in contact with some fibers
traveling at the rapid front roll speed and some at slower velocities.
The manner of drafting in one stage affects the performance of the next. Consequently,
there is a chain reaction along the stages of production, which can culminate in a very
poor performance of the spinning frame. Poor performance at this point directly
affects the efficiency of spinning and subsequent processes. It also adversely affects
the quality of the product.
A8.1.2 Error wavelength and amplitude
A machine-created error is often sinusoidal (i.e. its amplitude and wavelength are
characterized by a sine wave). An error wavelength is merely the distance along the
strand between repeats of the sine wave. Amplitude is the size of the maximum error
due to that sinusoid.
408
Appendix 8
A8.1.3 Mechanical errors
In practice, errors from the early stages of drafting are elongated by the drafting and
have long error wavelengths at the output of the final draft zone. Thus, there is a
spectrum of errors of varying wavelength, components of which come from the
various drafting zones. A repetitive error that is not sinusoidal can be expressed as a
Fourier series of sine waves of different wavelengths, and harmonic analysis can
reveal these various components. This is a valuable diagnostic tool for finding machine
errors and it is extensively used in the textile industry. Furthermore, random errors
typical of drafting waves caused by fiber-borne variations produce recognizable patterns
on a spectrogram. It might be recalled that the spectrogram is a diagram that gives the
error wavelengths and amplitudes.
Consider the effect of an eccentric front roll, where the roll is round but off-center.
Assume that the remaining rolls are perfectly true, all the rolls are of the same radius
(r), and the rotational speeds are ω1 and ω2 radians/sec for the back and front rolls
respectively. Let the eccentricity be ε inches. At the extreme position (a), the surface
speed of the roll is ω2(r + ε), but at the other extreme position (b), the surface speed
is ω2(r – ε) length units/sec. The linear velocity, Vo, of the delivery varies cyclically
between these extremes but the input velocity, Vi, remains constant; the result is that
the draft, ∆, varies cyclically. This, in turn, causes the linear density to vary. For an
eccentric front roll, the error wavelength is equal to the circumference of the faulty
roll. For cotton spinning this is often about 5 inches. (Bad aprons also produce cyclic
errors related to the length of the apron.)
In position (a), the mechanical draft is Vo /Vi = (ω2(r + ε))/(ω1r)
= (1 + (ε/r))∆
In position (b), the mechanical draft = [1 – (ε /r)]∆
Amplitude of the error = ± (ε/r) × 100%
[A8.1]
The mechanical error is magnified because meshing eccentric rolls have a nip line
that oscillates along the direction of the flow of fibers. This oscillation causes periodic
drafting wave activity and the physical movement of the nip itself creates an additional
error ‘spike’ in the spectrogram. If the faulty material is drafted again, the original
error is further elongated. Drafting reduces the absolute error amplitude but the
percentage value either remains constant or increases. The input error wavelength is
increased by drafting in proportion to the draft ratio. The general case is:
Error Wavelength = π × roll diameter × intervening draft
[A8.2]
Intervening draft is that draft which exists between the point of origin of the error and
the point of measurement of the strand. If there is no drafting between these points,
the intervening draft is 1.0. An error in the back roll of a drafting system produces an
error in the strand, but the error is then elongated as it passes through the draft zone.
The wavelength of the error in the material delivered by the front roll is equal to the
circumference of the back roll multiplied by the draft ratio between the back and
front rolls. Where several machines are involved, the appropriate draft is calculated
by multiplying all the intervening drafts together before applying the result to Equation
[A8.2].
Advanced topics VI: Drafting and doubling
409
A8.1.4 Fundamental theory of roller drafting
Much of the theory concerning roller drafting has treated individual fibers as long,
thin rods lying parallel to the flow direction, the length of the rods being the only
independent variable. The simple basis for the theory is that the fibers accelerate to
the front roll speed out of phase with one another. This causes migration of the short
fibers with respect to the longer ones [1]. Consider two fibers approaching a drafting
zone, as shown in Fig. A8.1(a), and consider a worst case scenario. The long fiber
approaching the input is of the same length as the roll setting and it is regarded as the
reference fiber. The short fiber is of length S and its leading end is level with the
reference fiber, of length L, as it passes into the back nip. The long reference fiber
cannot accelerate from velocity Vi to velocity Vo until the leading end reaches the
Draft zone
Reference
fiber
Vi
Vo
Short
fiber
S
L
L
(a) Time t = 0
Vi
Vo
V
(b) Time t = x
xi
Stage
1
Vo
Vi
U
2
3
4
l
L
xo
(c)
Fig. A8.1
a
Back nip Front nip
Floating fiber zone
Fiber flow in a drafting zone
410
Appendix 8
front nip, unless there is slippage between it and the roll surfaces. However, the short
fiber can accelerate before this and the acceleration point can vary. The short fiber
under these circumstances is called a ‘floating fiber’.
Figure A8.1(b) depicts the situation after the trailing end of the short fiber has left
the nip of the back roller. The fiber travels at a velocity V that is greater than Vi, and
could be as high as Vo, or somewhere in between (i.e. Vi > V > Vo). Fast moving fibers
already nipped by the front rolls but also in contact with the short fiber just described,
create a force tending to accelerate it (the short fiber). The more numerous fibers in
contact with the back rolls try to restrain this acceleration. The accelerating force
increases as the floating fiber moves towards the front rolls and the restraining
reaction decreases. Thus, there is a point at which the floating fiber accelerates to the
higher output velocity. From then on, until the leading end of the long reference fiber
reaches the front roll nip, the two fibers travel at different speeds.1 In the worst case,
the short fiber travels a distance (L – S) at velocity Vo after the acceleration. The long
fiber travels (L – S) at Vi, and the time for transit is t = (L – S)/Vi. Taking the draft
ratio, ∆, as numerically equal to the velocity ratio, the short fiber moves:
Vot = Vo(L – S)/Vi = ∆(L – S)
[A8.3]
The long fiber moves:
Vit = (L – S)
[A8.4]
Not all fibers migrate as much as this. The typical short fiber moves k(∆ – 1) (L – S)
relative to the reference fiber, where k is a factor < 1.
Grishin [2], like many others, assumed that the fibers are straight and oriented in
the direction of movement. Furthermore, he termed the longitudinal distance between
the centers of the fibers as shear; also he pointed out that the shear increases as the
fiber stream is drafted. Starting from the position of the leading ends of fiber, the
shear changes from xi (when both the short and long fibers are both gripped by the
back roll pair) to xo (when the long fiber passes from the control of one roll to the
next).
The scheme is shown in Fig. A8.1(c), where the long fibers of length L are shown
as heavy black lines and the short fibers of length l are shown cross-hatched. The
zone in which the short fiber can accelerate is shaded in gray and is called the
‘floating zone’. Fibers change velocity from Vi to U in the shaded zone. Most
theoreticians assume that U = Vo; the exact point where the events occur is open to
interpretation but clearly not all fibers accelerate at the same distance from the back
nip. Grishin assumed that xo = ∆xi and that the maximum deviation of a(∆ – 1)
occurred at the front nip. ∆ is the draft and a is the width of the floating zone. This
implies a uniform probability distribution of the fiber acceleration point over the
distance a. In this theory, a systematic displacement of all fibers of the same length
plays no role in the unevenness of the strand delivered. However, deviations from the
random do cause irregularity. Fujino and Kawabata [3] suggested that the wrongful
acceleration of the fibers in this zone is influenced by the fiber speeds. They deduced
that the probability distribution of the fiber acceleration point was an exponential
function of distance from the front nip. Goto et al. [4] used a normal distribution. All
1 According to some authorities, the short fiber passes through several steps of speed change,
perhaps sometimes accelerating during the intermediate stage, but, for the present purpose, a
simple one-step model will be used.
Advanced topics VI: Drafting and doubling
411
these theories assume that the fibers are straight, aligned and act independently.
Other authors recognize that fibers sometimes travel in groups. Whichever theory
might be closest to the truth, it is evident that the smaller the value of a, the less
chance there is for irregularity to develop. It is quite clear why a proper setting of the
aprons has such a beneficial effect on the evenness of the strand. The main limiting
factor in the use of aprons is the wear at high linear densities and high speeds. It is
also clear that the magnitude of the draft plays a large part in determining the errors
produced. Thus, one would expect that the ring frame would produce the largest
errors because the draft there is the largest of the roller drafting systems used in a
mill. Table A8.1 confirms this expectation.
Following Grishin’s lead, let the fiber shear be defined as the average distance
between the ends of various fibers denoted by X. Providing there has been no extra
disturbance, shear after perfect drafting is ∆X. If a fiber accelerates earlier than it
should, it becomes displaced relative to the others; it swims downstream because its
average velocity is increased by the early acceleration. The relative movement of this
fiber makes the strand thicker in the place to which it has swum, and thinner in the
place from whence it has come. Hence, the displacement of any fiber from its proper
position creates an irregularity of linear density. Such fiber migrations alter the shear
in the output. There is always some irregularity in the positions of the fiber ends and
the shear after drafting, X′, may be expressed as:
X ′ = ∆X + s
[A8.5]
where X is the shear, ∆ is the draft, and s is the standard deviation in shear. Shear may
be regarded as a form of longitudinal fiber migration.
Consider two fibers traveling at the same time, one of the correct length and one
shorter. The first travels through the draft zone at the back roll velocity until the
leading end almost reaches the nip of the front roll. The short one accelerates early
and travels distance y at the higher front roll velocity. The velocity difference during
the time that the short fiber is passing through the front nip is (V2 – V1) and the
relative draft is (∆ – 1).
The scale of all the deviations is changed by the same factor.
Standard deviation in shear = (∆ – 1)y
[A8.6]
Johnson [5] used a computer to simulate fiber movement in a random sliver where the
rolls were eccentric and the fibers were elastic. He also allowed the fibers to group.
Lamb [6] claimed that use of this mathematical model produced a good result for
wool processing. This demonstrates that, although the use of models using rigid rods
as the moving elements might be deficient, the models suggest that grouping of the
fibers might not be as important as some suggest.
Table A8.1
Typical changes in evenness of linear density due to drafting
Yarn count Ne
Theoretical CVth%
5% Uster values CVact%
Irregy index = CVact/CVth
Card
1st Draw
frame
2nd Draw
frame
Roving
frame
Ring
frame
0.14
0.65
2.7
4.2
0.15
0.67
2.8
4.2
0.18
0.74
2.8
3.8
2.0
2.45
5.2
2.1
60
13.4
14.9
1.1
412
Appendix 8
All the theories deal with the short-term errors, and even if some of these errors
are elongated by successive draftings, they deal with only one aspect of the problem.
A8.1.5 Drafting using toothed components (Staple fibers)
There is another class of drafting in which the one or more pairs of rollers is/are
replaced by moving toothed components. (A sketch of a typical toothed drafting
system is shown in Fig. A8.2(a).) These components may have saw-teeth, pins or even
just a roughened surface capable of gripping a fiber. The toothed components usually
create a grip on the leading portions of the fiber elements being drafted and the
trailing ends are often restrained by a roll and feed plate combination. Other combinations
are possible. In this class of drafting, it is necessary to define the fiber element being
drafted. The element might be a fiber clump of some size or it might be a single fiber.
Nose
Toothed roll
Feed roll
N
Vin
Feed plate
C
Vout
(a)
Fiber clump
F
F
Strength at weakest link = s
(b)
Front nip
Force
Back nip
Reaction force at N
for a strong strand CN
FS
Reaction force at N for
a weak strand CN
FW
Applied force
X
Y
Distance along the flow path
(c)
Fig. A8.2
Toothed drafting
Advanced topics VI: Drafting and doubling
413
Consider a sliver used in rotor spinning. There might be some very small fiber
clumps embedded in a matrix of a larger number of single fibers. However, the
average exit clump size nears that of a single fiber. In the opening line processes, the
clump size is much larger. These clumps are also embedded in a matrix of fibers and
it is difficult to characterize the clumps. Generally, the fibers within the clump have
a greater mutual cohesion than that which exists between them and those in the
surrounding matrix. The clump tends to retain its identity until sufficient force is
applied to break it apart. As before, the material enters at velocity Vi and is delivered
by the toothed element at Vo; the draft ratio is Vo/Vi.
Many theories of roller drafting relate to indivisible fiber elements, whereas with
toothed drafting the fiber elements (i.e. fiber clumps) divide in the process. Theories
of roller drafting assume that the length of element is virtually unchanged by the
drafting, whereas with toothed drafting it is almost certain that the average length
changes substantially. Consider a fiber clump being restrained by the rearward portions
of the mechanism (such as N in Fig. A8.2(a)). The teeth in the forward part of the
mechanism apply a force, F, to the leading portion of the fiber clump, which tends to
stretch the clump. This force increases as the leading end of the clump approaches the
virtual nip zone in the front of the drafting zone. As the clump moves forward, the
force rises until either (a) the trailing end is released, or (b) the clump divides at its
weakest point (Fig. A8.2(b)). One possibility is that, when the acceleration force, F,
equals the strength, s, at the weak spot, the frontmost portion of the tuft accelerates
to the delivery velocity. Later, a second weak spot might fail during the drafting
process. Then a further portion of the clump might accelerate to be followed by the
remainder when the new accelerating force and the reaction come into equilibrium.
Alternatively, the second weak spot may not fail before the equilibrium between the
accelerating force and reaction is reached, in which case the clump will have been
divided only into two pieces instead of three. Other possibilities also exist. The
statistical distribution of the daughter clump sizes differs from the mother distribution.
Generally, drafting reduces the range of clump lengths. The length of the clump is
always measured in the direction of flow.
Performance of the toothed system depends upon the design, among other things;
the shape of the zone between the toothed roll and the feed plate is particularly
important. An aggressive design tends to break down the clumps more quickly and to
straighten fibers more, but at an increased risk of fiber breakage. The design and
condition of the teeth or pins are also considerations.
In roller drafting, rule of thumb indicates that error wavelength of the major
drafting errors is about three fiber lengths. If this were applied to toothed drafting, it
would imply that the error wavelengths at the exit of the drafting system would be
about three clump lengths. As the tufts get shorter, so would the distances that the
tufts migrate due to the drafting process.
The greater the strength of the weak link, the smaller is the distance X in Fig.
A8.2(c), since the leading end of the tuft has to penetrate nearer to the nip to accumulate
a sufficient force. The strengths of the weak links that fail in drafting are variable and
the value of X is also variable. The value is equivalent to the value (L – S) in roller
drafting. It controls the migration of the daughter tuft with respect to a reference one
that does not divide. There is a distribution of fiber migrations and each component
contributes to the irregularity of the fiber stream emanating from the draft zone.
The complex and varying interactions between clump lengths, longitudinal migrations
and blending irregularities makes any deterministic solution extremely difficult. The
414
Appendix 8
distributions of these components are unknown and therefore only generalizations
can be made. Fiber migration is determined by draft ratios (in the order of 100). Thus,
it might be realized that relative fiber movements in the order of several yards are
obtained at each stage. After the various drafts of intervening machinery are taken
into account, cumulative relative motions in the order of 100 yards in card sliver are
possible. Such migrations might be significant when the blend is irregular.
A8.2
Roller drafting
A8.2.1 Draft distribution
The simplest practical roller drafting system is a ‘three-over-three’ system. Too small
a break draft does not fulfill its function of breaking down fiber clumps in the input
material and too great a break draft introduces unacceptable error in the first stage.
More attention to the quality of the input material pays a larger dividend than ultra
fine tuning of the break draft value. SKF [7] recommended that the break draft
should be between 1.1 and 1.4 for a total draft between 12 and 25; for higher total
drafts the break draft could be higher. Values up to 4 were suggested.
A8.2.2 Roll setting
There is always a margin allowed over the theoretical setting of the rolls. The exact
margin depends on the linear density of the strand and the variance in the length of
the fibers. For sliver, the setting can be as much as 10 mm more than the upper
quartile length (UQL) for cotton fibers. For finer strands, the margin is reduced.
However, other factors intervene and many machine manufacturers recommend settings
that are 1 or 2 mm less than the UQL. This is because natural fibers have only a small
percentage of long fibers; the majority of them are short. The breakage of a few long
fibers is regarded as a worthwhile sacrifice to obtain a better performance with the
bulk of the fiber population. Even with man-made fibers, there are fibers shorter than
the original cut length because of breakage and fiber convolution. Also, there are a
few that are longer due to stretching during processing. An example of the effect of
varying the setting for a drawframe when drawing polyester sliver is shown in Fig.
A8.3(c). The machine maker’s recommendation was for a setting of between 35 and
37 mm. The example was chosen because there is less variation in fiber length with
a man-made fiber than with cotton. When the setting was reduced below 27 mm, the
evenness of the sliver deteriorated sharply and it became progressively more difficult
to run the machine as the setting was further reduced. In theory, the relationship
between the setting and the fiber length is important. The settings theoretically should
be changed to match the fibers being run, otherwise the evenness of the output strand
suffers. However, in practice, it is rare for changes to be made in the mill except for
merge changes and during maintenance.
A8.2.3 Errors in drafting
Each draft zone produces its own error, which is superimposed on the errors already
present. Short-term periodic errors produce moiré effects, and cloudiness in a fabric
if the errors are non-periodic; long-term errors produce streakiness and barré.
In roller drafting, the top rolls are covered with a rubber material and, if the load
12
100% 38 mm
Polyester
Fiber
8
4
25
50
Roll setting (mm)
75
Roller neck motion
Uster CV, log scale (%)
Advanced topics VI: Drafting and doubling
C
A
B
20
30
(a)
40
50
60
Frequency (Hz)
(b)
70
180
Roll angle (degrees)
360
12
10
Roll separation
(arbitrary units)
CV of linear density (%)
415
8
6
4
0
1
2
Length along the strand (m)
0
(d)
(c)
Fig. A8.3
Some drafting errors
is left on while the machine is still, flats develop because the rubber is visco-elastic.
These cause periodic errors, which were a function of the ratch setting as in Fig.
A8.3(a).
It is common for top rolls to have several minor flats and Fig. A8.3(b) shows a
periodogram of the results found with a damaged top roll meshing with a true bottom
roll. The several peaks in the profile can be seen. When the data is expressed in CVs
as in Fig. A8.3(c), it is evident that the value of CV is considerably higher at start-up
than after a few seconds of running. Changes occur as the rubber of the top rolls
warms up. The variations in roll separation are closely tied to the irregularities in the
strand. A frequency of just over 30 Hz corresponds to the front roll speed (Peak A)
and Peak C corresponds to the second harmonic. Peak B comes from elsewhere.
When the separation of the rolls was plotted against angle of rotation of the bottom
roll for a very large number of rotations, there was a unique pattern dictated by the
roll error as illustrated in Fig. A8.3(d). Clearly, mechanical errors can be made to
show up distinctly, in contrast to the random errors (For more discussion see Section
A8.2.6). An investigation by Keyser et al. [8] showed a linear relationship between
yarn unevenness and roll eccentricity, as shown in Fig. A8.4. The data are old but the
result clearly demonstrates the importance of roll eccentricity. The authors also found
that yarn strength was diminished and the appearance of the yarn deteriorated. Foster
and Tyson [9] carried out a similar experiment and found that the slope of the curve
of standard deviation vs. roll setting was a function of draft. It follows that there is an
interaction between what happens in the draft zone and the cyclic change in velocities.
This aspect will be discussed later, but before that, other forms of error have to be
considered.
All significant mechanical errors in drafting arise from poor maintenance or improper
setting. Out-of-true rollers, rolls with uneven rubber hardness, and damaged roll
necks (i.e. bearings), gearing, aprons, or other components can cause product errors.
Slack bearings can also give problems. Cots or cushion rolls (rubber-covered top
416
Appendix 8
Unevenness (%)
140
Front bottom drafting roll
was made eccentric
y = 839x + 105
r2 = 0.846
130
120
110
100
0
0.01
0.02
0.03
0.04
Roll eccentricity (inches)
Fig. A8.4
Bottom roll eccentricity
rolls) must be buffed periodically to true them up; worn or damaged elements must
be replaced when necessary. The combination of a good maintenance program with
a proper quality control system to identify the sources of error is essential to the
running of a modern mill.
A8.2.4 Pneumafil and reworked fibers
Production of yarn temporarily ceases when an end breaks in a ring frame. An end
refers to the yarn emerging from the delivery rolls of the drafting system. Repair has
to wait until an operator (or robot) gets around to it. During these times of interrupted
production, the drafting system continues to deliver fiber, which is sucked away by
a pneumafil system. The fiber removed is also referred to as pneumafil. The fiber
removed is that which would have gone into the yarn, and has good characteristics of
length and strength; but fiber crimp and elongation characteristics have been changed
as compared to the virgin fiber. Nevertheless, the pneumafil waste is reused or
‘reworked’ because it is too good to throw away. The reworked fiber has to be blended
with the virgin fiber very carefully so that the percentage does not exceed, say, 3%
anywhere in the blend. Failure to control the blend leads to problems throughout the
process line.
Besides the pneumafil produced during the time of an end down, there is a continuous
loss of fiber from the twist triangle and balloon during spinning. This is only a
fraction of a percent of the total volume being processed, but it is significant in terms
of the amount of pneumafil produced. Consequently, there is not a unique relationship
between the ends down rate (in breaks/1000 spindle hour), machine productivity (in
lb/spindle hour), and the pneumafil production rate (in lb/spindle hour) as might be
expected, although there is a rough correspondence. Mill trials produce a range of
values depending on the product, machine, and operators. A very rough rule of thumb
is that the percentage of pneumafil is about one-tenth of the end-breakage rate.
Excessive deviations in the ratio between the pneumafil production rate and the
theoretical value given below are a sign of inefficient repairs of the broken ends.
Theoretical pneumafil production rate, Pp, is given by
Pp = K + {Py × (m r /ms)}
[A8.7]
where K = a factor dependent on the production of pneumafil created whilst the
spindle is working
Advanced topics VI: Drafting and doubling
Pp =
Py =
mr =
ms =
417
average production of pneumafil in lb/hr for 1 spindle
the production of yarn in lb/hr for 1 spindle
number of spindles idle in a set because of end-breakage
number of spindles in a set = (usually) the spinner’s assignment.
A8.2.5 Hairiness
Errors exist other than changes in linear density or creation of thick and thin spots,
slubs, and nep. If the surface of the yarn has a differing structure along its length, it
can produce customer complaints because of shading, barré, and moiré in the fabric.
Change in hairiness of the yarns is one of the factors responsible.
Variations in hairiness can come from several sources. One source is wear in yarn
guides and other running surfaces, which rough up the yarn as it runs through the
machine. Sometimes lack of careful maintenance will allow deep cuts to be produced
on surfaces, especially when running with a fairly abrasive fiber such as polyester. A
second source is in the twist triangle, where conditions leading to a ragged and
varying construction can also lead to undesirable changes in the surface of the yarn.
A third cause arises from varying quantities of short fiber arriving at the twist
triangle which produce changes in the yarn surface.
A8.2.6 Continuous measurement of sliver properties
Grover [10,11] discussed the problem of dynamic measurements of linear density in
a flowing sliver. Amongst many sorts of transducers assessed was a thermocouple
device (which measured the temperature of the throat of the trumpet), a pneumatic
system (which measured the pressure at the throat), and a force reaction system
(which measured the drag force acting on the active part of the trumpet).
Compressed sliver sliding through the throat of the trumpet dissipates energy.
Friction at the throat heats the sliver as well as the trumpet and it is possible to
measure the sliver temperature. The time response is slow for the relatively massive
machine components in the heat dissipation path. However, the speed of the sliver
causes it to respond to the changes in the resistance of the sliver passing through the
trumpet with a time response better than 0.05 sec. Significant differences in temperature
even over just a few inches of sliver running in a drawframe could be detected. When
running cotton, temperatures of up to 30°C were typical whereas with acrylic fiber
the value would rise to the region of 40°C. Drawn slivers produced lower temperatures
than carded ones. Figure A8.5 shows a series of scans with an infra red movie camera
recording the temperature of the sliver as it left the trumpet. Each scan showed an
increase in temperature as it passed over the sliver. About 70 mm of sliver passed
during one scan cycle. The three-dimensional graph shown has x, y, and z axes;
temperature is along the y axis, length along the sliver is along the x axis, and
distance across the sliver is along the z axis. All three axes are mutually perpendicular.
The temperature profile on each scan is an indication of structure of the sliver, and
the distance apart of the shoulders is an indication of the sliver diameter (as shown
in the diagram). Tests showed good results but the expense of the system was a
problem. Until a cheaper means of measuring the radiated heat is available, the
method is unlikely to be developed commercially.
Pneumatic trumpets are often used in which the air pressure at the throat of the
trumpet is measured as a proxy for linear density. The flowing fibers carry air into the
Appendix 8
Sliver temperature
at trumpet exit
418
y
30°C
26°C
0
Le
pa ng
ss th
ed of 1
th sliv
ro er
ug t
h ha
th t h
e
tru as j 2
m us
pe t
t
Distance
perpendicular
to strand axis
z
x
Sliver diameter
Fig. A8.5
Continuous sliver temperature measurement
trumpet and this creates a pressure in the throat. It also creates air turbulence due to
backflow at the entrance to the throat. The relationship between the air pressure and
the linear density depends on the bulkiness of the ingoing sliver and the speed at
which it travels. Trials were made with a trumpet in which a sliver was run and
pressure measurements were recorded. The sliver was cut into consecutive one-inch
lengths of sliver, weighed and matched to the corresponding portions of output signal
from the transducer. The comparison showed only a 0.64 correlation coefficient.
Longer-term errors yielded better results, but it was clear that the pneumatic trumpet
could not be relied upon to provide an accurate measure of very short-term error.
A trumpet was equipped with one or more diaphragms, each containing an orifice
through which the sliver passed. In the simplest one, a single orifice was mounted on
a diaphragm and strain gages were used to measure the reaction forces. A test with
consecutive one-inch samples, similar to the one just described, was used, and this
produced a correlation coefficient of 0.75. Although this was an improvement in
performance in the measurement of very short-term error, there were problems with
signal drift caused by uneven heating of the strain gage elements in a bridge network.
(A Wheatsone bridge measures and compares the electrical resistance of the four
arms; variations in strain or temperature of the material in any one of the arms will
unbalance the bridge and give a signal.) This emphasizes the difficulty of getting a
reliable signal to be used for control.
At a given roll pressure and a given fiber ribbon width, the linear density of the
sliver is given by the separation of the rolls through which the material travels. A
tongue-and-groove measuring system works on this principle and it is very successful
in yielding accurate results providing the operating surfaces are true and clean. However,
there is a tendency for the meshing of tongue and rolls to cut fibers that are not
pressed into the groove. Grover [10] and Lord and Govindaraj [12] used drafting roll
separation as a proxy for linear density of the ribbon of drawn slivers passing through
the front rolls. The measurements were found to be sufficiently accurate to use as a
control signal to adjust the ratch setting although some errors due to changes in
density at the selvages of the ribbon became evident. Mechanical errors in the rolls
came into sharp focus. The use of an encoder permitted the exact position of any flaw
in a rotating roll to be identified without stopping the frame. It was also found that
Advanced topics VI: Drafting and doubling
419
when a frame first starts up there are minor flats on the rubber cots but that these flats
disappear as the rubber warms up in use. Thus, there is a temporary increase in the
irregularity of the sliver when a machine is started after a certain rest time. The
regularity of the sliver improves over the first minute of running.
Another approach by the researchers just mentioned was to optically measure the
thickness of the fiber ribbon that passes through the draft zone of the drawframe.
There are difficulties in getting access to the material because of the space taken up
by the rolls. One of the conclusions from this phase of their work was that the input
slivers retained their autonomy through the drawing process, each sliver behaving
independently of the others. Traditionally, m input slivers are considered doubled
before drafting, and the variance of the material input to the drafting process is
expressed as (1/m)th of the average. The variance of the output material from the
drafting process is then given as:
(σout)2 = (σin)2 + (σadded)2
[A8.8]
where the variance due to drafting is added to the variance of the input material.
However, if the slivers behave independently during drafting, it is more correct to say
that the variance added by drafting is added to each input sliver and the doubling
takes place after drafting. The equivalent relationships are then:
m
m
m
j =1
j =1
j =1
(1/m ) Σ ( σ out, j ) 2 = (1/ m ) Σ ( σ in, j ) 2 + (1/m ) Σ ( σ added, j ) 2
[A8.9]
This may be interpreted as:
average of the output variances = average of the input variances
+ average of added variances
In this situation, drafting adds even more irregularities than is usually believed.
Furthermore, the result implies that their neighbors influence fibers when the packing
density is high, and that they behave as groups rather than as individual fibers.
Hence, the theoretical CVs should be calculated on the number of fiber groups in the
cross-section rather than the number of fibers.
A8.2.7 Automatic control of drafting errors
In 1962, Ishikawa and Shimuzu [13] proposed a device for sliver drawing in which
the drafting force was used to generate a signal to control the draft ratio. It was found
that the time constant of the system was an important parameter and that errors could
be amplified at wavelengths shorter than dictated by the time constant of the system.
In practice, they were able to control evenness down to about 3 inches (8 cm in the
original paper) but instabilities at smaller wavelengths were troublesome. They succeeded
in reducing the errors in their test apparatus but the importance of the paper was more
in directing attention to the regions of stability in a control system than in developing
a viable machine.
As mentioned previously, separation of the rolls in a drawframe has been used to
provide a signal, which was used to control the ratch setting rather than the draft ratio.
The idea was based on watching an expert set up a drawframe for minimum drafting
wave error. These errors are caused by variations in fiber attributes that react with
fixed roll settings. The expert first judges the drafting wave activity by the height of
420
Appendix 8
the hill on a spectrogram, then changes the roll setting, makes another test, re-judges
and so on. This process continues until the setting is optimized for minimum drafting
wave activity.
In the automatic control system, the middle and rear pairs of drafting rolls were
mounted on a platform, the rolls were driven by stepping motors, and the setting of
the front drafting zone was also controlled by a stepping motor. A computer program
was written that sampled the output from the roll separation transducer within a
specified error waveband, which included a typical drafting wave. If the magnitude
of the signal was outside a control zone, the ratch setting was adjusted by 1 mm and
the error was sampled again. If the error was still out of control, the process was
repeated, but if it was now in control, the process ceased until the signal went outside
the control limits again. Roll errors made sharp peaks enclosing only a small error in
a periodogram whereas the drafting waves made a diffuse spectrum that encompassed
a considerable variance. The system performed satisfactorily and could be seen to
change settings from time to time as the fiber population changed. A cotton sliver
was spliced to one made of longer polyester fibers and, as the splice passed through
the system, there was a permanent adjustment to the new fiber length. Despite this,
the system was not successful because it was so slow in reacting to changes.
A8.3
Avalanches in roller drafting
A8.3.1 Slub production
As stated earlier, a short fiber in the draft zone not being nipped by either set of
rollers is called a floating fiber. The floating fiber is under the influence of others in
contact with the front and back nips but it is not directly controlled by either set of
rolls. Consider what happens just before and after the acceleration of the floating
fiber.
Just before the acceleration, the subject fiber may be in contact with other floating
fibers. For explanation purposes, assume the following. Three fibers in contact with
the subject floating fiber are completely controlled by the back nip. Two are adjacent
floating fibers traveling at the speed of the back nip but they are not directly connected
to it. Another fiber is completely controlled by the front nip and is also in contact
with the subject. No other fiber controlled by the front roll is in contact with the
subject. If, at that moment, the fiber in contact with the front roll generates not quite
enough force to overcome the frictional resistance between the subject and those in
contact with the back roll, the subject will not yet accelerate. If, however, one of the
three which had been in contact with the back roll is later released, and accelerates,
it might cause the subject also to accelerate because there is now one fiber less
restraining it. If the subject does accelerate, there is a possibility that the other
floating fibers in the vicinity will be induced to go with it. A small avalanche has
been created. Obviously, larger avalanches are possible, especially where the subject
fiber and neighbors are nearing their normal acceleration points anyway. Where fiber
clumps rather than single fibers are involved, the greater number of neighboring
fibers can have a greater effect because the circumference of the clump is larger than
that of a single fiber. Consequently there are differing chances of avalanches. The
effect of such avalanches is to magnify the size of the defects and to undesirably
increase the variability of output.
If a group of fibers accelerates early, then the population of fibers in the rearward
Advanced topics VI: Drafting and doubling
421
section is depleted. Also, the group to be accelerated is not formed until Z inches of
strand have passed. After drafting this becomes ∆Z. Concentrations of short fiber
create weak zones, which are interspersed with good fiber. The body of the avalanche
can become a slub, which carries a concentration of short fiber with it when it
accelerates. These concentrations of short fiber are likely to cause more avalanches
(and slubs) in later processing. After drafting, there are sometimes long portions of
good material between the outbreaks of slubs (which are often processed without
trouble). Under these conditions, there can be outbreaks of this type of fault in each
of the drafting systems through which the concentrations pass. A steady concentration
of short fiber can break up into pulses of short fiber; these, in turn, break up in
subsequent drafting into intermittent bursts of activity. Relatively short lengths of
yarn may be subject to heavy production of slubs and then there can be quite long
periods without activity. The final slubs can be very heavy and the following thin
spots are very prone to breakage in spinning. To minimize such effects, the roll
settings must be properly controlled and high concentrations of short fiber should be
avoided.
A8.3.2 Variations in fiber population
Lord et al. [14] point out that a fiber (or fiber clump) accelerates when the force
exerted on it by fibers being pulled into the front nip exceeds the reaction from the
more slowly moving fibers under the influence of the rear gripping medium. In roller
drafting, this would be the back roll. The acceleration point is dictated not only by the
relative speeds of the fibers, but also by the number of fibers in contact with the
subject fiber (clump) and their degree of entanglement. This may lead to unstable
conditions and avalanches [15].
Table A8.1 makes it clear that the actual CVs of linear density are much larger than
the theoretical values for the draw- and roving frames where the strand cross-sections
are large. It is not until the cross-section is reduced to the order of 100 fibers that the
theoretical and empirical values approach one another. This suggests that the fiber
flow in the first stages of the process line is not composed of independent fibers
moving singly in relationship to one another; rather that fibers move in groups, and
this causes a deterioration in the expectation of evenness.
It also suggests that the populations of fibers entering a machine may vary
considerably along their length. If this is so, then there will be corresponding variations
in drafting performance. Lord and Johnson [15] suggested a system of fiber population
changes that would carry errors from one machine to the next in line, to make
complex patterns of spinning performance. Grover [16] made many cross-sections of
polyester/cotton yarns. The percentages of each sort of fiber were noted at one-inch
intervals. The work gave CVs of about 20% in polyester content. The few gaps were
filled by interpolation. The resulting time series (shown graphically in Fig. A8.6(a))
was subject to a fast Fourier transformation to give the periodogram shown in Fig.
A8.6(b). The CV in the one component was considerably higher than that of the
linear density (c 14%). There seemed to be a distinct probability that the main
variability detected was from drafting waves in the roving frame. Since then, some
work in measuring the characteristics of the cotton fiber delivered from the drafting
system of a ring frame has shown a CV of linear density over 17% in long-term
variability. The high CV probably arose from variations in the first drawframe. It is
very difficult to measure the short-term variability except for short lengths. The
422
Appendix 8
40/1 P/C yarn
Amplitude (arbitrary scale)
Polyester in X section (%)
40
35
30
25
20
15
500
FFT of % polyester
400
300
200
100
0
10
0
25
50
75
100
Length along the yarn (inches)
(a)
Fig. A8.6
1
10
100
Wavelength, log scale (inches)
(b)
Variation in polyester content
implication of this work is that changes in fiber population interact with the successive
drafting zones to cause varying performance. The characteristics of batches of yarn
might well be associated with these changes in fiber population. Changes in the
attributes of the fiber passing through the system undoubtedly affect the CV of linear
density of a yarn emerging from the front rolls of a ring frame. These effects might
be even larger than those arising from changes in roll settings or machine condition.
We finish this section on a practical note. Sampling the production and testing the
CV gives a sample value that is good only for that time. A number of samples taken
over an extended period needed to characterize a drafting system.
A8.4
Doubling associated with roller drafting
A8.4.1 Reduction of variance by doubling
Associated with many drafting systems is the idea of doubling. Several strands of
input material are fed to the drafting system in an attempt to reduce the errors. A
typical machine in which this occurs is the drawframe. Here it is not practical to use
aprons; only the fixed control surfaces and the effects of doubling mitigate the irregularity
caused by drafting. A quite usual assumption is that the variance in linear density
among the strands is averaged. The resulting variance of the output is taken as (1/m)th
of the variance of any of the m slivers used in the creel as input (i.e. m doublings).
This neglects the variance added by drafting and the variations between the input
slivers. An idea of this can be gleaned from an old study by Ozgur [17]. Data were
re-plotted to give Fig. A8.7. Plotting the output irregularity in terms of 1√m and
extrapolating towards zero, the irregularity does not reduce to zero. In fact, some
studies have found that drawing more than about three times not only reduces sliver
cohesion but also causes the regularity to deteriorate. Also, Bowles and Davies [18]
showed that the improvement in CV due to doubling is reduced as the wavelength of
error increases. No improvement will occur if the input error wavelength is much
larger than the length of sliver in the supply cans. Very long variations can be induced
by a poor bale laydown and doubling at drawing or elsewhere after carding might
have little effect on these longer components.
Dyson [19] considered a model in which the variability of fiber extent was introduced.
Advanced topics VI: Drafting and doubling
3
2
2.5
1
0
Sliver irregularity (CV%)
5
y = 6.861x + 0.777
r 2 = 0.986
r = correlation
coefficient
Draft ratio = 8
Doubling = 8
Sliver irregularity (U%)
4
423
0
0.2
0.3
0.4
1/√Doublings
NB Irregularity is a synonym for unevenness.
0
0.1
Fig. A8.7
Effect of doubling
(Fiber extent is the distance between the extremities of a crimped fiber.) The minimum
irregularity, CVi, is then expressed as:
CVi = {100/( mk )}(1 + 0.0001 CVd2 + 0.0001 CVk2 )
[A8.10]
Where CVd is the CV of fiber linear density, CVk is the CV of fiber extent, m is the
number of fibers in the cross-section, and k is a factor. This assumes that fiber
fineness and fiber extent vary independently of one other.
Dyson quoted work by other authors yielding k = 0.95 for carded cotton ring spun
yarns and k = 0.8 for rotor yarns. With k = 0.95 and 0.8, the equations simplify to
109/√m and 119/√m, respectively.
One way of expressing irregularity is by using an index, which relates the actual
CV to the theoretical value. Experience shows that the index of irregularity decreases
as the material moves down the process line, with the index varying from roughly 4
at the card to about 1.1 at the ring frame. Lamb [6] argues that doubling does not
affect the index because both the denominator and numerator are affected by the
factor √m. Undoubtedly, the movement of aggregations of fiber rather than single
ones has an effect.
A8.4.2 Effects of between-stream variance
Passing reference was made earlier regarding the need for similarity in the means and
variance in the streams of material being doubled. An example in staple yarn production
illustrates how the between-stream variance can be included in the estimate of the
expected total variance or CV.
Figure A8.8 illustrates how variations in the means of strands A, B, and C broaden
the probability distribution of the combined strand shown as D. The means are a, b,
and c; and the total unevenness in the product stream is calculated by adding variances
between the means to the variance within the samples. In symbols:
CVt ≈ √ ( CV 2 / m + CVb2 )/100%
[A8.11]
where CVb is between the means, CV is the mean of the components, and m is the
424
Appendix 8
a
Statistical frequency
CVa
A
Components
CVc
CVb
C
B
b
CVt
c
Total
D
O
Linear density
Fig. A8.8
Variability caused by differences in means
number of strands. Doubling of periodic errors can present problems, particularly if
the error frequencies are similar for all input strands. The relative longitudinal positions
of the streams then play a part in determining the error. Any such difficulties in this
respect can be avoided by making sure that the equipment is in good condition and
does not produce periodic error.
A normal drawframe has input slivers of varying mean values of linear density.
Thus, finisher drawn slivers provide variations in yarn not only from the variance
within each sliver and the variance produced by spinning but also from the variance
between the slivers.
A8.4.3 Doubling mass constant
As discussed elsewhere, the flow of material through a mill is not really continuous;
rather batches of material are processed in sequence. When traditional doubling is
used, it is within a batch. Thus, for example, when we double sliver in a drawframe,
we double within the batch defined by the mass of fiber in the creel. This mass may
be thought of as a mass constant. The system is not able to significantly reduce error
for wavelengths greater than that represented by about twice the mass constant of the
machine involved.
A8.4.4 Effects of overdrawing
Whilst drawing sliver improves the orientation of the fibers and applies some doubling,
too great a number of drawframe passages can adversely affect the product. Sliver
tenacity falls off rapidly with multiple drawings. Klubowicz [20] determined the
effect of multiple drawings (up to 36 drawings) on the yarn strength and strand
evenness. He found that the yarn strength reached an optimum at somewhere between
four and eight drawings, but then decreased with further drawings. The strand uniformity
and yarn appearance improved with increased number of drawings, while the yarn
elongation decreased. There was difficulty in handling overdrawn sliver because of
the low sliver cohesion. This makes it clear that there is a limit to the benefits derived
from drawing and doubling. Improvements in fiber orientation are similarly limited.
Advanced topics VI: Drafting and doubling
425
Many of the simple ideas of doubling and drafting are insufficient to explain the
whole set of problems.
A8.4.5 The combined effects of drafting, doubling, and twisting
Cavaney and Foster [21] found, from empirical studies, that the variance of the output
strand from a drafting system was:
(ANe(∆ – 1)/m] + b
[A8.12]
The factor A was a figure of merit but was not a constant and the factor b was almost
zero; m was the number of ends fed to the system. Speed of the frame was found to
have little effect. They recognized that fibers did not necessarily travel through the
draft zone independently; they commented that the variance depended, in part, on the
number of fibers in the fiber groups involved. Further, they pointed out that twist in
a strand had a stabilizing effect on drafting. The performance of a roving frame may
also be expected to differ from that of a drawframe on that account.
A8.5
Doubling and toothed drafting
A8.5.1 Opening line and carding
The process of dividing fiber clumps, which is a form of drafting, was described in
some detail in Chapter 5. As outlined in Section A8.1.5, the basic ideas are fairly
clear but the idea has not been widely recognized. There are large drafts applied to
fiber clumps and the division of the clumps is irregular. Draft is applied to the whole
stream, and the flow becomes irregular. However, few attempts have been made to
assess this irregularity because (a) it is difficult to do so without impeding the
operation, and (b) the effect of the irregularity is obscured by the massive doubling
that occurs in devices like mixers and chute feeds. Perhaps someone will realize that
there are potential gains in better controlling the fiber flow in the opening line, and
then we shall see a further step in the continuing trend of improved yarn quality.
A8.5.2 Rotor spinning
Separation of fibers in a strand supplied to an open-end spinning machine is the
essence of the process. It is necessary to separate the fibers almost into separate
entities to make the system work. Open-end spinning, in the form of rotor spinning,
has become very successful. Toothed drafting is an essential part of that success.
Damage to the feed and combing rolls produces periodic errors as one might suspect
and that sort of error can be detected by conventional testing and corrected by proper
maintenance. Random variation in the fiber stream delivered to the rotor is reduced
by the massive doubling that occurs when the many layers of fiber are laid inside the
rotor to build up the necessary linear density of yarn (see Sections 3.4.1 and 7.2).
Short-term random errors are low in rotor spinning. Perhaps the main lesson to be
learned is that adequate doubling reduces the random errors. For this to be effective
the mass constant must be large enough and, of course, there must be a sufficient
number of doublings.
426
Appendix 8
References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
Lord, P R and Grover, G. Roller Drafting, Text Prog, 23, 4, 1993.
Grishin, P F. A Theory of Drafting and its Practical Applications, J Text Inst, 45, T167, 1954.
Fujino, K and Kawabata, J. Method of Analyzing Problems in Drafting, J Text Mach Soc
Japan, 8, 3, 1962.
Goto, H, Ichino, S and Kurozaki, J. Fiber Motion in Roller Drafting, J Text Inst, 48, T389,
1957.
Johnson, N A G. A Computer Simulation of Drafting, J Text Inst, 72, 2, 69, 1981.
Lamb, P R., The Effect of Spinning Draft on Irregularity and Faults, Parts I and II, J Text Inst
78, 88 and 101, 1987.
SKF Bulletin No 3, Skefko Ball Bearing Co Ltd, Luton, UK, Technical notes, 1957.
Keyser, W R, Middleton, J O and Dougherty, J E. The Effect of Roll Run-out in Spinning on
Yarn Quality, Text Res J, 1956.
Foster, G A R and Tyson, A. The Amplitudes of Periodic Variations Caused by Eccentric Top
Drafting Rollers and their Effect on Yarn Strength, Trans Text Inst, T385–T393, 1956.
Grover, G. Dynamic Measurement of Sliver Properties, Ph D Thesis, North Carolina State
Univ, Raleigh, USA, 1989.
Grover, G and Lord, P R. Measurement of Sliver Properties on the Drawframe, J Text Inst,
1992.
Lord, P R and Govindaraj, M. Dynamic Measurement of Mechanical Errors in Sliver-Drawing,
J Text Inst, 81, 195, 1990.
Ishikawa, S and Shimuzu, J. Automatic Control of Short Term Sliver Irregularities by the
Detection of Drafting Force, Text Mach Soc Japan, 8, 3, July 1962.
Lord, P R Stuckey, W C, Yu, X and Grover, G. Deblending in Roller Drafting, J Text Inst, 76,
5, 339, 1985.
Lord, P R and Johnson, R H. Short Fibers and Quality Control, J Text Inst, 3, 145, 1985.
Grover, G A. Periodic Variations in some Fiber Properties along Strands Drafted in a Roller
System, MS Thesis, NC State Univ, USA, 1984.
Ozgur, N. Relationship between Input and Output Variations at Drawing, MS Thesis, NC.
State Univ, 1965.
Bowles, A H and Davies, I. Description of the Evening Action at the Drawframe. Part III,
Extremely Long Wave Irregularity, Shirley Inst Bull, 42, 1, 19, 1969.
Dyson, E. Some Observations on Yarn Irregularity, J Text Inst, 65, 215, 1974.
Klubowicz, A MS Thesis NC. State Univ, USA, 1965.
Cavaney, B and Foster G A R. The Irregularity of Materials Drafted on Cotton Spinning
Machinery and its Dependence on Draft, Doubling and Roller Setting, J Text Inst, 46, 8,
1955.
Appendix 9
Advanced topics VII: Yarn balloon mechanics
A9.1
General observations
The whirling length of yarn between the pigtail guide and the bobbin produces yarn
tension. Too high a tension above the pigtail guide leads to a high frequency of endbreaks, which reduces spinning efficiency and yarn quality. Too high a tension below
the traveler makes unwinding at the next process stage more difficult and increases
the number of interventions in winding, which reduces both winding efficiency and
yarn quality. Too low a yarn tension in the balloon leads to collapse that produces
similarly undesirable results. The behavior is usually analyzed by considering the
forces involved, but there is also the possibility of using energy balances as a means
of description. Forces are vector quantities whereas energy is a scalar quantity and
this provides some relief from the mathematical rigor needed for acceptable solutions.
An explanation can be derived from a consideration of the energy dissipated, stored,
and/or transformed at the various parts of the rotating system. We will look at both
approaches.
Technicians observe the balloon shape as a measure of the yarn tensions. Sophisticated
means of measurement are rarely available in a mill and the normal means of judgment
is whether the balloon is long and thin, whether it is fat, or whether it is bottle shaped.
What is being observed is the surface area swept by the rotating yarn. Theoreticians
have used vector mechanics to explain the complex phenomena and a reference point
frequently used is the node formed at the pigtail guide. However, the energy to
sustain the balloon derives from the bobbin, which is rotated by a mechanical drive
system.
As far as the present discussions are concerned, the subject will be divided into
several divisions. These will deal with various aspects of ballooning, and then go on
to discuss (a) the lower zone between the traveler and the winding point, (b) the
central zone, and (c) the upper portion above the pigtail guide. The central zone is a
powerful tension producer but it is the reactions at the traveler and pigtail guide that
produce the consequences.
428
A9.2
Appendix 9
A rotating plane balloon – a very simplified case
As matters of definition, let the surface swept by the generator OBP (Fig. A9.1) be
called the balloon and that swept by the more or less horizontal generator QP be
called the base. To introduce the ideas, first consider the problem at the lower of two
levels of simplification; the simplifying assumptions reduce the obscurity of the
topic without enormously affecting the central idea. In this simplified case, the yarn
rotates in a vacuum about an axis OQ, as shown in Fig. A9.1. This involves a rather
unrealistic assumption that the plane OPQ rotates about the axis OQ at a speed of ω
radians/second and that the yarn is confined to this plane. The theory ignores the
effects of bending and torsional stiffness of the yarn. The yarn is treated as a string
of beads, each element of which is δs units long. The action of the distributed
centrifugal force and the restraints at O and P cause the yarn to become curved
similarly to a hanging cable. Several radii may be defined. The radius with respect to
the spindle axis at any height is denoted by r. The maximum value (re) is at the
equator at B. The radius rg is that at which the mass of the yarn length OBP may be
considered to be concentrated (i.e. the distance to the centroid G). The length of yarn
in the balloon (S) above the traveler may be estimated by calculating the length of the
line OBP. The mass of yarn in that length is Sn (where n is the linear density of the
yarn) and the centripetal acceleration acts on that mass through the centroid G.
During a single chase, the length S might change by up to 10% but, in this rough
analysis, the change will be ignored. (The meaning of the word chase is defined in
Section A9.3.2.) Ft and Fho are the horizontal components of the forces acting on the
yarn at the traveler and pigtail guide, respectively. Ft includes the centrifugal force
acting on the traveler. The centrifugal force Fcf = nω2 rgS.
Horizontal force components must balance and the moments within the plane
should also balance. Taking moments about O,
Fcf × hg = Ft × H
[A9.1]
Fho
O
hg
H
Ft
ω
G
B
Q
Fcf
P
rc
Fig. A9.1
Rotating plane
Advanced topics VII: Yarn balloon mechanics
429
Ft is equal and opposite to the difference between the horizontal components of
centrifugal force acting on the traveler and the reaction of the traveler with the ring.
Distribution of the forces between O and P also alters with the position of the
centroid, which changes during the building of the bobbin. The centrifugal force
acting on the yarn changes with the radius of the centroid, rg. The effect of these
important reactions will be described later. A force of Fcf acting on an element δS is
ω2rnδS. The total force acting on the portion OP is the summation of all such elemental
forces between O and P. If the length of the yarn between these limits is S, then
s
Fcf = Σ ω 2 r n δ S
[A9.2]
o
This force is divided as in Equation [A9.1] to give the reactions at the ends. The
vertical components of force for the simplified model are shown in Fig. A9.2. A force
acting along the yarn is called yarn tension, T, and a force in any other direction is
denoted by F. In a portion of the yarn at the equator B, the downward tension is Fv
and the upward reaction at O is Fvo. For vertical equilibrium, Fv = Fvo.
A similar argument can be made if the yarn had been cut at U. The vertical
component of T at that point also equals Fv for equilibrium. Thus, in the oversimplified
model, the vertical component of tension in the yarn balloon does not vary at all as
a function of height; it is solely determined by the end conditions. In ring spinning,
the forces at the traveler and the pigtail guide determine the end conditions. The
resultant forces at these points press the yarn against metal and friction creates
tension gradients across these items. The frictional forces cause the tension below the
pigtail guide to be more than that just above it because the yarn moves downward.
The tension in the yarn leaving the traveler is higher than that just above it because
there is a tension gradient due to friction. These effects will be further discussed later.
These rough analyses set the scene and establish how the distribution of the applied
force alters the reactions at the end points. Even when the balloon is considered in
Fvo
To
O
hg
φ
φ
Fcf
H
x
B
U
Z
Ft
y
Q
Horizontal
components
(a)
Fig. A9.2
O
Yarn
tension
Fho
To
re
Fv
Q
Vertical
components
(b)
Force components in a simple theoretical balloon
430
Appendix 9
three rather than two dimensions this is still true, although the analysis becomes more
complex.
A9.3
Energy distribution in the balloon
A9.3.1 Energy taken from the bobbin
Figure A9.3(a) shows a compound view of the yarn between the bobbin and the
ωy
B
Yarn
Bi
ωt
ωb
(a)
Bo
Rail traverse
Bobbin
Traveler
Ring
αo
ωt
Bo
Bi
αt
O
A
ωb
rBo
rr
rBi
Plan view
(b)
T1
Fy
T2
A
Yarn
traveler
Fa
B
Ring flange
Reaction
View perpendicular to
plane containing the yarn
(c)
Fig. A9.3
The lower portion of the balloon
Advanced topics VII: Yarn balloon mechanics
431
traveler. The torque supplied to the rotating yarn system is the mathematical product
of the horizontal component of yarn tension at B and the winding radius. The energy
supplied is torque × rotational speed.
Let this energy be designated E. Conservation of energy dictates that the energy
available at B is absorbed or dissipated by (a) kinetic energy of the yarn between the
pigtail guide and the point B in Fig. A9.3, (b) kinetic energy of the rotating traveler,
(c) strain energy stored in the yarn under tension and torque, (d) losses due to airdrag
acting on the yarn, (e) friction losses caused by the traveler sliding on the ring, and
(f) energy arising from forces generated by balloon instability. Potential energy changes
are negligible; energy changes due to balloon instability are ignored.
A9.3.2 Energy balance in the base
The first component to consider is the more or less horizontal portion of yarn lying
between the bobbin surface and the traveler. Figure A9.3 shows oblique and plan
views of the ring and traveler system. The plan view at the bottom shows only the
yarn departing from the traveler on its way to the bobbin and does not show the yarn
arriving. This is to make it clear that the angles between the center line and the yarn
change. As the winding point reciprocates between the lay point on the bare bobbin
at Bi and that of the full bobbin radius at Bo, the angle changes from αi to αo. Due to
this motion, vector components of the yarn tension acting along OA vary from TBo sin
αo to TBi sin αi, as the chase moves from bottom to top. (The chase describes the
reciprocating movement of the lay point of the yarn onto the conical portion of the
yarn already on the bobbin. The lay point, or wind point, means the point on the
bobbin surface where the yarn is laid.) Periodic changes in geometry of the yarn, as
the winding point moves through the chase, are reflected in the yarn tensions.
Kinetic energy stored in this portion of yarn is:
Ek1 = I1 ω2/2
[A9.3]
where I1 is the second moment of mass of yarn in the base about the axis of the
bobbin and ω is the rotational speed in radians/second. The subscript ‘k’ refers to
kinetic energy and ‘1’ refers to the base.
The winding radius changes cyclically through B0 and B1 as the ring rail moves
through the chase motions. Further, the length of yarn changes cyclically through
AB0 and AB1 (Fig. A9.3(b)); consequently, the kinetic energy in AB changes cyclically
because of alterations in length, mass, radius of gyration, and winding radius. Ek1 is
a factor that depends on the geometry and mechanical arrangements of the short-term
ring rail motion (or chase). The torque available changes cyclically in sympathy with
the rail movement.
The kinetic energy of the traveler Ekt = Itω2/2, where It is the second moment of
mass about the spindle axis. Since I t = M t k t2 , where Mt = mass of the traveler, and
kt is its radius of gyration (note: radius of gyration is a special term used in mechanics
to describe not only the position of the mass, but also its shape and size). The
subscript ‘t’ refers to the traveler and ‘kt’ to the kinetic energy of the traveler.
E kt = M t k t2 ω 2 / 2
[A9.4]
The elongational strain energy is Tε/2, where T is the yarn tension, and ε is the
elongation of the length of yarn involved. Yarn is visco-elastic and thus there is also
a non-recoverable energy loss associated with the extension which is proportional to
432
Appendix 9
the length of the yarn segment concerned. However, the length of yarn in the segment
now being considered alters. Consequently, the strain energy (Es1) changes in sympathy.
Thus, it follows that the energy level has a component, which is affected by the chase
motion, but the changes are small and may be ignored.
Among other factors, airdrag depends on the length of yarn in the airflow. The fact
of changes in length of AB means that there is a cyclic change in airdrag on the
particular segment of yarn that is synchronous with the chase movement. Consequently,
the energy dissipated (Ea1) has a dependence on the chase similar to those just
discussed. The subscript ‘a’ refers to airdrag. Airdrag losses in this segment of yarn
are minor compared to the kinetic component; there is no need to complicate the
analysis further. There is also a yarn–metal frictional energy loss at the traveler, but
the sliding velocity is so low that this item, too, may be neglected.
The total energy absorption between A and B (which we may denote as E1 where
E1 = Ea1 + Es1 + Ek1) has a cyclic component that is dependent on the chase motion.
Thus, the energy available to the traveler and the yarn above it is E – E1. It might be
realized that E is a variable by virtue of the changes in yarn tension and, as has just
been discussed, that E1 has a component related to the cyclic chase motion.
A9.3.3 Friction between ring and traveler
The traveler is pulled round the ring by the yarn. Drag on the traveler due to the
friction between it and the ring causes the yarn in the balloon to rotate slower than the
bobbin. The difference in speed causes the yarn to ‘wind on’ the bobbin. Referring to
Fig. A9.3, the relative rotational speed of the bobbin in relation to the traveler is (ωb
– ωt), and the relative linear winding speed is (ωb – ωt)rb. The rotational speeds of the
yarn and traveler are the same except for occasional local excursions in portions of
the yarn above the traveler. The fiber is delivered to the system at constant linear
speed and the winding system adjusts itself accordingly. Movement of yarn along its
own axis proceeds at constant velocity, Vy. Values of ωb and Vy are fixed, but rb
changes with the position of the winding point within the chase. Thus, the rotational
speed of the yarn, ωy, changes with ring rail position within the chase but it is
normally marginally less than the bobbin speed. Microwelds between the ring and
traveler, unstable air conditions, and perhaps some other causes result in occasional
deviations from the normal cycle of events.
Figure A9.3(c) shows portions of the traveler and the sliding track on the ring; the
traveler is shown cut at the level of yarn contact for clarity. Components of the yarn
tensions in contact with the traveler (T1 and T2 in Fig. A9.3(c)) produce a resultant Fy.
The tension at A is not the same as at B. A component of Fy tends to hold the traveler
away from the ring. Centrifugal force acting on the traveler (Fct) acts horizontally
along a radius centered on the spindle axis and it tends to force the traveler into
harder contact with the sliding track on the ring (i.e. in a direction roughly opposite
Fy). Forces Fy, Fcf, and Fa adjust themselves to provide equilibrium by causing the
traveler to tilt as necessary. The force normal to the sliding track, Fn, is the vector
sum of the appropriate components of these and the tangential friction force along the
sliding track is µFr (see Section A9.5.5 for the full analysis of forces.) The energy
dissipated is Eft = µFr × ωrr, where the subscript ‘t’ refers to the traveler and ‘f ’ to
friction. Since ωrr is virtually constant, the energy loss depends on the coefficient of
friction, µ, and the normal force, Fn. The coefficient of friction is affected by the
lubrication, or lack thereof; for cotton, lubrication is largely from crushed fiber
debris deposited on the track on the ring.
Advanced topics VII: Yarn balloon mechanics
433
The coefficient of friction depends on the state of wear of the sliding surfaces and
the reaction force depends on the attitude of the traveler, yarn tensions, and centrifugal
forces. Wear on the traveler alters the position of the center of contact area on the
traveler. Also, the magnitude and directions of the yarn tension vectors vary. This is
important because, not only do the forces applied to the traveler have to balance, but
so do the first moments. Thus, if the direction and magnitude of Fy alters, the traveler
tilts to correct the imbalance. This results in a modified value of Fn and a change in
the energy absorbed in friction. In other words, there are reactions to changes in the
system both above and below the traveler. There is also a long-term variation caused
by wear in the components. In the case of the traveler after the initial break-in, this
long-term change is measured in days, whereas the corresponding change to a properly
run-in ring is measured in months, or even years.
The kinetic energy of the traveler, Ekt, = Itω2/2 = M t k t2 ω 2 / 2 , where the second
moment of mass about the spindle axis = It, the mass = Mt, and the radius of gyration
of the traveler = kt. Thus, the mass of the traveler is an important factor in determining
the yarn tension since ktω varies but little.
Thus, summarizing:
Eft = µFn × ωrr
[A9.5]
E kt = M t k t2 ω 2 / 2
[A9.6]
and
A9.3.4 Yarn above the pigtail guide
Rotation of the balloon induces torsion in the yarn above the pigtail guide and this
stores some energy as torsional strain energy; it dissipates some of this due to frictional
losses. There is also a small amount of tensile strain energy involved. The strain
energies Es reduce the energy available to the main balloon, but the quantity involved
is small and may be neglected.
A9.3.5 Energy available to the main balloon
The energy available to the yarn above the traveler is (E – E1 – Ekt – Eft – Es). Changes
in the chase motion, the mass of the traveler, and the effects of wear are now seen to
affect the energy available to the main yarn balloon. As before, the energy available
is distributed over categories similar to those already recited. Kinetic energy of the
upper yarn depends on the mass of yarn involved and its radius of gyration; strain
energy depends on the length of yarn involved. Of these factors, kinetic energy is the
most important and the integration implicit in the factors for airdrag1 may be left for
later. Let Ea2 be the airdrag of the yarn in the balloon between the pigtail guide and
the traveler. The subscript ‘2’ refers to the main balloon. The kinetic energy now
available to the main balloon is:
(E – E1 – Ekt – Eft – Es – Ea2) = Ek2
[A9.7]
1 Airdrag on an element of yarn may be taken as proportional as to Cd α √n (ωr)2δ, where Cd is
the airdrag coefficient, α takes into account the attitude of the yarn element, n is the linear
density of the yarn, ωr is that linear velocity which is tangential to the circular locus of the yarn
element and δ is the length of the yarn element.
434
Appendix 9
Let Ek2 be the kinetic energy, n the linear density, S the length of yarn, and k the
radius of gyration; each term referring to the yarn in the balloon rotating about the
spindle axis, and the yarn referred to is between the pigtail guide and the traveler.
Equation [A9.1] may be modified as:
Ek2 = nSk2 ω2/2
[A9.8]
Where Ek2 is the kinetic energy, n is the linear density, S is the length of yarn,
and k is the radius of gyration; each term referring to the yarn in the balloon
rotating about the spindle axis. The yarn referred to is between the pigtail guide and
point B.
The radius of gyration is related to the maximum diameter of the balloon; the
normal speed is assumed to be so slightly different from the spindle speed that it can
be regarded as constant. Thus, if ω and n are treated as invariable, the changes in
energy available must cause changes in S, or k, or both. In other words, the size and
shape of the balloon changes with the energy available. The length, S, changes
significantly as the yarn on the bobbin builds up from base to tip; also there are
changes in k. The yarn spirals in the balloon depending on the airdrag and this can
cause significant changes in length; there is no unique relationship between length
and diameter. These changes in Ek2 are superimposed upon those arising from the
right-hand side of Equation [A9.7]. Whatever combinations of these factors exist,
there is a change in energy level that is distributed over the time it takes to spin a
bobbin full of yarn.
A9.3.6 Instabilities
It is possible for the traveler motion to become unstable. At these times, the attitude
of the traveler oscillates and imposes an additional energy variation on the main
balloon, which is reflected in the tension variations.
If the yarn tension at the wind point drops below a certain level, there is insufficient
energy available to maintain the normal, single-noded balloon. The balloon will then
change such that the length of yarn is accommodated in a different shape, which
permits the radius of gyration (or, approximately, the distance of the centroid) to be
reduced. Sometimes the change in shape involves wrapping part of the yarn around
a revolving support; when this happens the unsupported length and the radii of
gyration are reduced. There are then extra frictional losses. The balloon is said to
collapse because of the reduced diameter. One cause of such an event is the use of a
traveler of too low a mass. This will be discussed later.
It is possible that subsidiary oscillations in the main balloon could absorb energy
that would be subtracted from that available to shape the mean path of the yarn. There
could be resonant vibrations at surprisingly low frequencies because the yarn in the
balloon is restrained at the ends like a suspended cable; the effective modulus of
elasticity of the system is low. Energy losses due to such vibrations would be one of
these subtractions. Vibrations of this sort might be excited by intermittent slippages
of torque and tension at the pigtail guide, mechanical vibrations, local air disturbances,
etc. The system has a response time to force pulses imposed upon the balloon. Microwelds between the ring and traveler can create such pulses, particularly if the ring is
not properly run in. Recovery of normal running conditions after such occurrences is
dependent on the response time.
Advanced topics VII: Yarn balloon mechanics
A9.4
435
Yarn tension gradients
A9.4.1 Tension gradients as a connecting factor
A factor that appears in nearly all the energy items mentioned in the previous sections
is yarn tension. There is a progressive change in yarn tension from the bobbin to the
fiber delivery system, which is situated above the pigtail guide. A change in one
segment affects all the rest; thus, the factors discussed earlier are mutually dependent
and some discussion about yarn tension is necessary.
A9.4.2 Dynamic and passive yarn tension gradients
Tension gradients in the yarn may be classified into two categories. One has been
called dynamic, because it arises from the centripetal accelerations acting on the
yarn. The other has been called passive, because it does not depend directly on the
rotation of the yarn about the spindle axis. Figure A9.4 illustrates one case and
demonstrates how yarn tension forms a connecting thread in the control loop of the
system. Remembering that the winding tension helps determine the available energy
for the system, it can be seen that the behavior of any one segment of yarn is
dependent on the tensions generated elsewhere.
Yarn tension is at its highest at the winding point on the bobbin. The existence of
a dynamic tension on the more or less horizontal portion of yarn sweeping the base
means that the tension at the traveler is less than the maximum. Since the outer
diameter of the balloon base is limited by the ring radius, this gradient does not vary
much for a given spindle speed. Friction between the yarn and the traveler causes a
passive tension gradient and the tension of the yarn entering the traveler is still lower.
The yarn in the main balloon suffers dynamic tension gradients that vary along the
length. A passive, frictionally induced tension gradient occurs at the pigtail guide
with the result that the input tension to the twisting section is further reduced. In Fig.
A9.4, the curve is shown for the simplest case, but the gradient can be multi-noded.
When the balloon collapses, the lower end of the yarn might wrap around the bobbin
(or crown, if one is used) and a further passive tension gradient will be introduced.
It is also possible for the balloon generator to change from a roughly parabolic shape
Length along the yarn
Pigtail guide
Tension gradients
due to frictional
contact with the
guides
Yarn flow
Traveler
Wind point
Yarn tension
Fig. A9.4
Yarn tension profile
436
Appendix 9
into a sinuous one with a number of nodes. In some sorts of unwinding from a
bobbin, the balloon base, ring, and traveler no longer exist and these sources of
tension gradient are removed from consideration.
A9.5
The real balloon
A9.5.1 The central section of the real balloon
When operating without balloon control surfaces (more about these later), any element
of yarn is subject to the tensions and forces acting on it. The forces include airdrag,
other frictional restraints, and the effects of electrical charging. Few of the forces
arising from these phenomena act through a common point and thus there are moments
that tilt, bend, and twist the yarn in the vicinity of the element. Figure A9.5 illustrates
the forces acting on an element of yarn. The yarn in the balloon is curved; consequently
the tensions acting on an element of yarn do not act along a common straight line;
furthermore, the tensions differ from end to end. The forces acting approximately
within a horizontal plane are shown with cross-hatched arrows. The light gray area
represents the horizontal plane.
Let us take the roughly horizontal forces one by one. The airdrag is due to the
relative motion between the yarn element and the surrounding air. It does not act
along the same line as the velocity vector because of the airflows caused by the
pumping action of the bobbin and yarn [1]. The centrifugal force acts along a line in
the horizontal plane that passes through the center of rotation of the yarn element.
This may or may not be congruent with the center of rotation of the bobbin. The
Yarn tension (T + dT )
Fm
Fa
Mechanical drag
Air
drag
Fcf
Yarn
Centrifugal
force
Horizontal plane
Yarn
Velocity V
Yarn tension T
Fig. A9.5
The central zone of a balloon
Advanced topics VII: Yarn balloon mechanics
437
mechanical drag force, Fm, arises when the hairs protruding from the yarn element
lash some machine part such as a separator plate, or when there is shear in the airflow.
This produces minor periodic, false twist torques as repeated contact is made. The
force system comprises components from the cross-hatched force vectors shown. The
yarn element tilts and twists to balance the system, as sketched in Fig. A9.6.
Effects of tension, twist, and mechanical abrasion can change the condition of the
yarn and alter the airdrag characteristics and perhaps the propensity to react to mechanical
disturbances, such as contact with separator plates. The magnitude of these various
components varies from one level to another because of the changes in radius, linear
speed, airflow, and physical condition of the yarn. The shape of the yarn in the
balloon is curved in three-dimensional space; also the gradients of tension, torque,
and geometric attitude vary from level to level within the balloon. The varying
population of fibers in the yarn being spun produces varying linear densities and
amounts of hairiness, and there are long-term variations in airdrag that alter the
tension patterns in the balloon [2]. These, of course, influence the end-breakage rates.
A real balloon is not confined to a rotating plane as was earlier assumed. The
effects just discussed cause it to spiral, and if we use a rotating plane in any mathematical
model, we can use it only as a reference. Thus, the real balloon rarely fits the simple
theories; comprehensive equations of motion are needed. These give a better fit but
they are not easy to manipulate without a computer and even these sophisticated
programs do not completely account for all the vagaries of the balloon.
The next step is to consider events with such a reference. Plane OPTr in Fig.
A9.6(a) is assumed to rotate at the speed of the traveler about OP, and yarn in the
balloon does not necessarily even touch it. (Tr is used rather than T to avoid confusion
with tension.) Yarn streams behind the winding point due to the frictional forces and
any element above the traveler lags. For example, an element at B lags the traveler by
φ, with the yarn taking up an angle α, to a vertical plane OPBQ that passes through
the segment being considered. Some drag may be from mechanical friction with
Plane OPQ
O
Wind and
drive point
E
Fv +
Torquei
ω
α
Fh +
B
Yarn
B
Fh
Tr
Airdrag = Fa
P
α
φ
Traveler
Q
TorqueO
(a)
Fig. A9.6
Fv
A non-plane balloon
(b)
438
Appendix 9
machine components, and there is an airflow caused by the moving parts. However,
for the moment, we will deal only with airdrag caused by the motion of yarn through
a fixed environment. Clearly there are three components in directions: (a) tangential
to the locus of the element concerned, (b) vertical and parallel to OP, and (c) horizontal
and parallel to QP.
A9.5.2 Airdrag
Airdrag is a function of fluid friction caused by the yarn moving relative to the air.
McAdams [3] quotes the Fanning equation, which indicates that drag is proportional
to the square of the relative linear velocity and is a complex function of Reynolds
Number.2 Figure A9.7 is based on McAdams’ data which refers to flow in pipes but
is often used for fluids flowing outside, but parallel to, the pipe axis. Our case
involves hairy yarn and the hairs stream behind the main body almost like a comet’s
tail and affect the coefficient of airdrag. The relative airflow is usually neither parallel
nor perpendicular to the comet’s tail. The tail is oriented away from the line of motion
because the hairs are subject to centrifugal as well as airdrag forces.
In calculating Reynolds Number for airplane wing sections and the like, the typical
dimension normally used is the chordal width or length of the streamer rather than
the thickness. Other researchers use results from flow perpendicular to the yarn, and
the resulting graph has a somewhat similar shape but a different scale. The ‘comet
tail’ of fibers is thought to have a significant effect on airdrag in ballooning. Some of
the yarn near the top operates in the laminar region, with high drag coefficients.
Other parts operate near the equator in the turbulent region with lower drag coefficients,
and intermediate parts operate in the unstable region with variable drag coefficients.
The point of the diagram is not the friction factor, but the range of operating conditions
involved and the instability around 103 to 104 Reynolds Number. Fortunately, the
highest drag coefficients occur at small radii and thus have only a small effect. Figure
A9.8 demonstrates differences in the lag of the yarn due to airdrag. Not only do the
theories give differing results for the portion of yarn below the equator, but experimental
data show that considerable variation is possible. Most theorists assume the yarn to
1.0
Friction factor
Usual assumption
Unstable
0.1
0.01
Laminar
0.001 2
10
103
Fig. A9.7
Turbulent
104
105
Reynolds Number
106
Airdrag coefficient
2 Reynolds Number = ρVD/ζ, where ρ = air density, V = relative velocity, D is a typical dimension,
and ζ is the viscosity of the air. It is the ratio of viscous and inertia forces; at a critical value,
the flow changes from streamline to turbulent.
Advanced topics VII: Yarn balloon mechanics
Stationary
r = –3/52h2 + 4.54h – 0.06
Non-stationary r = –3.77h2 + 4.63h – 0.03
Normalized height (h)
0
ω
270°
240°
Node
210°
0.25
Re
E
180°
0.50
Equator
300°
Non-stationary =
Stationary
=
Experimental =
0°
O
B
Φ
0.75
150°
0
0.5
1
1.5
Normalized radius (r)
(a) Elevation
Fig. A9.8
30°
A
Ring
1.00
439
60°
120°
90°
(b) Plan view
Radius profile of a balloon
be a thin cylinder of yarn of diameter d, and calculate Reynolds Number and drag
coefficient (Cd) accordingly. If the length of the hairs streaming behind the yarn is
used rather than the yarn diameter, the Reynolds Number spans a range that includes
laminar, unstable, and turbulent regimes.
A9.5.3 Balloon theory relating to the central section
Batra et al. [4,5] quote the basic equations of motion of a quasi-stationary balloon,
the adjective ‘stationary’ referring to the fact that they assumed the yarn was stationary
relative to a rotating plane of reference. The plane of reference included the axis of
rotation of the bobbin, and rotated about that axis. A balloon node was at the pigtail
guide situated on the axis, and the center of the ring was also situated on the axis. The
yarn in the balloon was treated as a quasi-stationary object with respect to the traveler
and then the mathematical description was boiled down to a series of differential
equations amenable to solution. The vector equation is:
Absolute acceleration = A0 + Ar + 2ω × vr + ω × (ω × s) + a × s
[A9.9]
Some of these terms can be eliminated and the following comments apply: (a) the
acceleration of the origin at the pigtail guide is zero, (b) the first term (A0) is zero, (c)
the second term (Ar) is negligibly small, (d) the third term containing the Coriolis
acceleration is negligible, and (e) the last term is assumed to be zero because it
includes the factor a, which is the angular acceleration of the balloon. Thus, the
equation reduces to absolute acceleration = ω × (ω × s), which can be translated as
centripetal acceleration acting along the radius of rotation of an element = –ω2r.
Furthermore, the rate of tension change along the rotating yarn is –nω2rdr, where n
is the linear density of the element, ω is the rotational speed in radians per second, r
is the radius of the element, and dr is the incremental change in radius over the
element considered. From this, Batra et al. say that the tangential component of airdrag is negligible and that the tension in the yarn in the balloon is:
To – T = nω2r2/2
[A9.10]
As the yarn is made heavier, the balloon enlarges, the spindle is run at a higher speed,
440
Appendix 9
the traveler weight increases, or any combination of them, the yarn tensions increase.
A curiosity is the resemblance to the equation met in rotor spinning (centrifugal force
= n ω2r2/2); the yarn inside the rotor rotates within a plane and in a balloon it
occupies a three-dimensional space.
The validity of using a stationary model is disputed by Lisini et al. [6] for cases
where the balloon shape is subject to rapid variations. They point out that in ring
spinning, the movement of the ring rail causes changes in traveler speed and this
undermines the assumption of constant rotational speed of the inertial frame. The coil
of yarn deposited on the bobbin forms a spiral rather than a circle. Relative motion
between the traveler and the wind-on point is caused by changes in length and
attitude of the yarn between the wind-on point and the traveler. These changes are
related to the alterations in length in the yarn forming the balloon above the traveler.
Consequently, the traveler speed tends to change cyclically, but the variation is small.
Thus, it is true that an error is involved in using the traveler to anchor the inertial
frame even if the effect is small. These authors favor the finite element method of
calculation over the iterative Runge-Kutta solution. (The finite element theory assumes
that the yarn is made up of very small straight segments.) A comparison of the two
methods shows that they give similar results above the equator of the balloon, as
shown in Fig. A9.8(a). However, the plan views in the top left quadrant of Fig.
A9.8(a) make more visible the differences between the two theories relating to the
yarn lying below the equator.
It is interesting to note that the shape of the elevation of the balloon is very near
to parabolic, confirming the data of the present author. The angular lag of elements
in the balloon relative to the traveler varies in the two theoretical cases. The use of
stationary solutions greatly simplifies the analysis but at the cost of some accuracy.
Theoretical models involve non-linear equations and a computer is required to obtain
a solution in a reasonable time span. However, the solution is only as good as the
assumptions made in respect of airdrag, coefficients of friction, and the flow of
torque and tension in the system [2].
At point B back in Fig. A9.6, there is a system of forces that includes those shown,
but there are others perpendicular to the plane of the paper. Curvature of the element
results from the application of these forces. The normal to the yarn at B no longer
intersects OP. The center of curvature is in space outside the balloon. Tension gradients
across the yarn segment make the forces at the upper terminal of the segment differ
from those at the bottom one; the segment is forced to tilt until the moments are in
equilibrium.
At different heights above the traveler, the inclination of the yarn, α, changes with
respect to the center line. The element of yarn shown does not lie in the plane of the
paper but at an angle that varies. Yarn tension is the resultant of all the components
acting on a segment terminal and as α alters, the yarn tension changes. There are
sometimes multiple solutions to the equations under unstable spinning conditions.
Instability is often the result of the use of too light a traveler.
Figure A9.8 also shows a plan view of some yarns in a balloon. As previously
mentioned, the top left quadrant contains theoretical data based on the work of Lisini
et al. [6]. An adjustment was made to the angular positions of the stationary and nonstationary curves to bring them as nearly as possible into congruence. It is interesting
to note that there is little difference between the results for that part of the yarn that
lies above the equator (which normally includes the majority of the yarn in the
balloon). In the bottom right quadrant there are two sets of new experimental data
Advanced topics VII: Yarn balloon mechanics
441
gathered within a few seconds of each other, with the spinning machine running at
constant speed and a fixed rail height. The data in the two quadrants should not be
compared because different conditions prevailed, however curves A and B in the
bottom right quadrant should be nearly identical but they are not. Obviously, variations
in the yarn altered the shape of the balloon. One candidate for suspicion is the airdrag
coefficient, which is normally modeled as a constant for a given yarn.
A.9.5.4 Balloon control rings
The purpose of a balloon control ring (see Fig. A9.9) is to reduce yarn tension; the
device works for a range of conditions but it is not universally effective. Balloon
control rings cannot be effective under the conditions of incipient collapse and they
are rarely used for fine counts. From a practical point of view, the surfaces can
become poisoned by accumulations of fiber finish or oligomer and these accumulations
lead to difficulties in spinning. The control rings also impose an extra drag on the
yarn that increases the spirality with effects similar to those discussed above. The
control rings also tend to make the yarn more hairy.
As a first step, one can use a fairly superficial explanation of their mode of
operation. The control rings reduce the surface area of the balloon. When the spiral
angle of the yarn in the balloon is small, the yarn tension is roughly proportional to
the surface area of the balloon. Thus, the control rings pinch the balloon to form a
waist, which reduces surface area and thereby reduces the yarn tension. Mathematical
models confirm that control rings reduce the tension for stable balloons and promote
stability; the rings also reduce the destabilizing influence of slubs passing through
the balloon.
The reduction in yarn tension permits the use of higher speeds, weaker yarns, or
both. A higher speed improves productivity and permits the spinner to spread the
fixed costs over a larger poundage, which reduces the cost/lb. The possibility of using
Node
β
Control ring
movement
Drive
Main ring
movement
Yarn removal
Fig. A9.9
Main ring not shown
Balloon control ring
442
Appendix 9
weaker yarns means that, sometimes, lower twist can be used and this also increases
productivity. Advantages are balanced by disadvantages. Summarizing the problems
with balloon control rings: (a) they make the yarns more hairy, (b) they accumulate
spin finish, (c) they add slightly to the cost of the machine, (d) they interfere with the
doffing and piecing operations, and (e) they produce a torque in the yarn within the
balloon. Of these, the first two are the most important.
A9.5.5 The traveler
The balloon size and shape vary as the yarn builds up on the bobbin, and this is
associated with changes in yarn tension. Consider the forces acting on the traveler as
depicted in Fig. A9.10. Centrifugal force, Fct, acts through the center of gravity of the
traveler and is balanced by the resultant yarn tension, Fy, also there is a reaction force,
Fr, acting between the ring and the traveler. There is a sliding contact between the ring
and traveler at A, and the friction due to this exerts a drag force which causes the
traveler to lag behind the bobbin. The beauty of the system is that the speeds adjust
automatically to the prevailing conditions; no mechanical complications are needed.
Sliding contact can cause serious wear on the traveler and the life of the traveler
is then measured in days. A normal practice is to judge the wear of the travelers by
the number that are burned. According to Grishin [7], every 10% of burned travelers
in the population increases the ends down rate by 5 per 1000 spindle hours. There is
also some collateral damage to the ring and, over a much longer time, the ring too
becomes unserviceable. For no damage, the vector sum of Fy, Fr, and Fct should be
zero, but if we were to run under those conditions there would be traveler instability.
Centrifugal
force acting
on traveler
Fct
Resultant yarn
tension Fy
H
Traveler
Yarn
Yarn
T2
A
Flange
Centroid
of traveler
x
Ring Reaction force F
r
cross section
Tilted
traveler
Flange
Flange
damage
x
T1
Traveler scar
View in direction x – x
(b)
(a)
Yarn
Y
Fy
ωt
Ring
X
Fr
Fct
Traveler
Bobbin
ωb
Fcd
(c)
Fig. A9.10
Ring and traveler
Advanced topics VII: Yarn balloon mechanics
443
Fy varies during the bobbin build and, if contact is to be maintained, the traveler
weight has to be sufficient to control the tension over the whole range of conditions.
The reaction force, Fr, is strongly influenced by the traveler mass, M, and consequently
the tension, TB, is also dependent on it. Traveler mass has to be changed as the yarn
count is altered within the normal spinning range (M/n is usually kept constant).
Adjustments also have to be made for changes in ring size and shape.
A properly run-in ring will last for years, whereas a traveler might only last, say,
10 days. The coefficient of friction between the two metal surfaces changes and this
influences the drag force. The tensions T1 and T2 (Fig. A.9.10(a)) cause the traveler
to tilt and the angle of tilt changes with balloon geometry. The reaction force is
sufficient to cause transient metal to metal seizures of the poorly lubricated surfaces,
although fiber debris and particles of fiber finish offer some lubrication. As a new
traveler is put into use, there is a small contact area that runs at a high local temperature
and creates fairly rapid wear of the surfaces. The damaged surface of the traveler is
concentrated in a band and Fig. A9.10(a) shows a scar typical of a used traveler. Wear
causes the area of contact to increase sharply at first, but the rate of wear then abates
as the scar on the traveler grows. As the scar widens, the centroid of the reaction
moves and changes the attitude of the traveler. Eventually, the tilt becomes sufficient
to cause it to be thrown off. Before that happens, however, the yarn tensions become
sufficient to cause a higher end-breakage rate than normal. It is important to change
the travelers in timely fashion.
If moments are taken about A in Fig. A9.10(b), the moment due to the resultant
yarn tension must balance the moment due to the centrifugal force acting on the
traveler. Any change in the geometry of the traveler alters the position of the centroid
and causes the traveler to adjust its angle with respect to the horizontal until balance
is achieved. Various factors determine the forces described. For a given ring diameter,
the centrifugal force acting on the traveler is determined by its mass and speed. The
linear density of the yarn, the balloon geometry, the rotational speed of the balloon,
and the reaction between the ring and traveler define the resultant yarn tension. For
a given speed, the centrifugal force acting on the traveler is theoretically constant
whereas the forces transmitted by the yarn vary as the bobbin builds. Thus, the angle
of tilt taken up by the traveler varies cyclically. With a poor design of traveler, slipstick conditions can lead to an unstable porpoising as it rides the flange and this
either causes an end-break or throws off the traveler.
Because of constraints in mass and size of the traveler, there is little space available
for the yarn and it could become trapped near H. Consequently, the shape of the
traveler is important and each type of yarn not only needs a traveler that has the
required mass, but one which provides adequate space for the yarn. The yarn can also
become trapped if the traveler tilts too much. If the traveler is too heavy, the friction
between the ring and traveler soon destroys the traveler and might damage the ring.
If it is too light, the balloon can collapse and cause high tensions with all the problems
described earlier. Stability of the yarn package becomes a problem if the winding
tension (related to traveler weight) is reduced too much; soft-wound packages occupy
too much volume and are liable to become damaged in subsequent handling.
Frazer [8] illustrated the instability of a balloon when the traveler is too light. At
low traveler mass, there is an ambiguous tension at radius r, as indicated by the
leftmost dark curve in Fig. A9.11. In that case, three tensions are theoretically possible
at the lowest traveler weight shown. The other dark curves show stable relationships
between the lay point radius and tension.
Appendix 9
Yarn tension
444
r
Lay po
int radiu
s
Fig. A9.11
v
Tra
r
ele
ma
ss
Effect of traveler mass
A9.5.6 The lower portion of the balloon
The factors determining the resultant force (Fy) can be visualized in three dimensions
as indicated in Fig. A9.10(c). It will be seen that Fy is the vector sum of the forces in
the upward pointing section of yarn at Y and the roughly horizontal section shown at
X. The resultant force is balanced by the system containing: (a) centrifugal force
acting on the traveler (Fct), (b) the reaction between traveler and ring (Fr), and (c) the
drag force, Fd, acting tangentially to the ring. The drag force (Fd) is the result of the
traveler sliding on the ring at a rotational speed of ωt. The bobbin (shown in truncated
form) rotates at ωb. Reiterating previous statements, changing the mass of the traveler
alters the yarn tensions. Indeed this is the only practical way a user can adjust the
tension, given that the yarn count, balloon geometry, speed, and machine configuration
are fixed by design or commercial considerations. The centrifugal force acting on the
traveler is Fct = ω2rtM, where M = mass of the traveler, rt = radius of the locus of its
centroid, and ω = rotational speed. The angles taken up by these portions of yarn are
critical in determining the tensions TA and TB, which act at X and Y respectively.
These tensions, in turn, help to determine the rest of the tensions in the balloon. There
is a relationship between them that is dictated by the friction forces between the yarn
and the traveler. Using Amonton’s Law as an approximation, the relationship is:
TA = TB eµΨ
[A9.11]
However, any assumption that the coefficient of friction is a constant is imperfect.
The coefficient varies with yarn hairiness, finish, and possibly rh. Another source of
possible error in Equation [A9.11] arises from the fact that the yarn is bent to a small
radius of curvature when passing round the traveler. If the radius is too small, bending
stiffness begins to play a significant part and the normal force between the yarn and
the traveler will be higher than estimated (which results in higher drag force). Thus,
one might expect deviations in winding tension from those predicted by some
mathematical models. Also, the greater the energy loss in overcoming the drag force
acting on the traveler, the less is the energy available to inflate the balloon.
During spinning, the yarn winding point is controlled by the ring rail motion.
There is a fairly short oscillation period as individual cones of yarn are laid on the
bobbin. Also, there is a much longer period as the bobbin is built from bottom to top,
laying new cones over each of the previous ones. Each chase builds a new layer of
Advanced topics VII: Yarn balloon mechanics
445
yarn and requires a small change in mean height of the ring rail. The upwards rate of
change of the rail position during the chase is usually different from the downward
one; this is to create an interlocking yarn package structure.
A vector component of the tension along ABo in Fig. A9.3 helps to balance the
centrifugal force acting on the traveler. Let this force be TB cos α and let Fd = kbµ Fr
(where kb is a factor to take into account the forces omitted and Fr is the reaction to
the forces acting on the traveler). The appropriate subscripts should be added. If
the yarn above the traveler is nearly upright, k is almost 1.0. Substituting for Fd and
cos α = rwo/r, we can write in functional form:
Fr/TB = f{(rw/rr), kb, µ}
[A9.12]
The radius rw varies from rB1 to rB0; k and µ vary also. This relationship implies that
as the ring rail moves, Fr /TB changes. The tension TB is related to the yarn tension
above the traveler and it increases to a local maximum at the top of the chase where
rw becomes a minimum. There is a limiting size to the bobbin diameter. As stated
earlier, the bobbin size is normally about 40% of the ring size, because winding yarn
on smaller diameter bobbins creates excessive yarn tension. Also, the bobbins are
slightly tapered. As the bobbin builds, the wind-on radius, rw, is normally limited
between about 0.4rr and about 0.9rr.
For continuous control, Fr > 0 if undesirable instability is to be avoided. The
energy available to the system = Twrwω, where the subscript w refers to conditions at
the winding point. Except under the unlikely condition where tension Twrw is invariable
and ω changes significantly, any changes in rw are associated with changes in energy
available. Because of the changes in yarn angles at the traveler, the passive tension
gradients can change markedly with changes in rw. Under stable conditions, an increase
in rw is associated with a drop both in tension and energy available. The tension
variations arising from the ring rail movement, which controls the chase, are measured
at frequencies of less than 1 Hz.
A9.5.7 The upper zone of the balloon
Yarn tension between the node at the pigtail guide and the front rolls of the drafting
system is a critical factor in determining end-breakage rates in spinning. The friction
of the yarn running through the pigtail guide situated at O in Fig. A9.12 affects the
tension and twist of the yarn. As mentioned earlier, To ≈ Tieµε. If the guide is offcenter, or the yarn flow approaching the guide is not coaxial with the center line of
the spindle, the angle ε varies within each revolution of the yarn in the balloon, with
the result that Ti varies also. The point at which the strand is at its weakest usually lies
in the twist triangle and Ti must be kept below that breaking strength. The term strand
can mean either the yarn or the fiber flowing through the twist triangle. Not only is
the value of To important, as previously discussed, but so is the angle ε because, if the
variation is large, then ω2rr has to be kept lower to compensate (ry is the yarn radius
in general). The tension variations from this source appear at the frequency of the
rotation of the balloon (say, 200–300 Hz).
Not only is there a tension gradient in the yarn passing through the pigtail guide,
but there is also a torque gradient. The normal force acting on the yarn at the contact
point produces a friction drag force, which has components (a) along the yarn, and
(b) tangential to a normal cross-section of yarn. The former produces tension and the
latter produces torque. It can be argued that:
446
Appendix 9
Tension Ti
Yarn flow
O
ε2
ε1
Tension T01
Fig. A9.12
τo ≈ τieµα
Tension T02
Upper portion of the balloon
[A9.13]
The symbol τ denotes torque and the other symbols have their previous meaning. If,
for example, we assume that µε varies between the limits of 0.04 to 0.12 radians as
the package builds up, then the ratio of torques would vary between 1.04 and 1.128.
In other words, the average twist in the yarn above the pigtail guide would, in that
case, be reduced by 4% to 13% compared to the value just below the guide. The twist
is reduced in the very place where it might be an advantage to increase it. Attempts
to use rotating guides to overcome the problem have not been successful; this is
partly due to the extra costs involved and partly to the difficulties in piecing.
False twist in the yarn leaving the pigtail guide is of some importance. The resultant
of the input and output yarn tensions on either side of the guide has a horizontal
component that presses the yarn against the inside surface of the pigtail. The yarn
might roll, as well as slip, on that surface and the rolling action would produce false
twist between the drafting system and the pigtail guide. Total twist in this region is
the sum of the real and false twists but the false twist above the guide is negative. The
net effect is another small reduction in twist in the yarn coming away from the twist
triangle. This component changes cyclically.
The twist triangle geometry is determined by the net twist, which affects a number
of yarn properties such as hairiness and bulk as well as the end-breakage rate. If there
are surges of twist at this point, then the balloon will be disturbed, the tension will
fluctuate, and the yarn properties will vary accordingly. The surges are similar to
those sometimes found in rotor spinning at the navel.
Evidence of such phenomena has been gathered by illuminating the balloon with
Advanced topics VII: Yarn balloon mechanics
447
horizontal thin sheets of light at different levels. The normal assumptions imply that
the loci of the small segment of yarn are circular. In some cases seen in industry, the
locus of the yarn elements just below the pigtail guide is badly distorted. Figure
A9.13 shows a modest distortion arising from the pigtail guide, but it fades at distances
remote from the guide. There are also distortions from other causes. The strata designated
B through E were between the pigtail guide and the top of the bobbin. The stratum A
was above the guide in the secondary balloon and the stratum F was below the top of
the bobbin. The balloon control ring did not operate and the photographs were taken
with the ring rail at a constant position in the chase.
A9.5.8 Stability of the speed of the yarn balloon
There is a torque generated tending to change the rotational speed of the mass of an
independent element of yarn, δm, if it changes radius. As a first model, consider a
yarn to consist of a series of contiguous elements like a string of beads and the length
of yarn in the balloon, S, is ∑δm. If elements of yarn do not follow a circular locus,
they must change speed to conserve momentum unless a pattern of forces restrains
the change. Momentum of each element = Iω and I = δm × r2 = nδs × r2, where n is
the linear density of the element. In a balloon, the elements of yarn are not independent,
but a change of radius still produces a system of forces tending to change the speed
of the element and of its neighbors. A reduction of radius causes the elements to
speed up and an increase in radius slows them down. Another cause of change arises
from the lag of one element relative to another due to drag. Any change in drag alters
the transient speed of the element with respect to the lower portions of yarn in the
balloon. Once the stability of the balloon is disturbed, transient changes in speed and
shape of the balloon are inevitable. These effects are not normally large unless the
balloon is in or near the unstable region.
A
Pigtail guide
B
C
D
E
F
Fig. A9.13
Loci of balloon elements
448
A9.6
Appendix 9
Balloon collapse
A9.6.1 Energy variation in the balloon
The relationships between the forces acting in a balloon are complex and distinctly
non-linear. Most often, a perturbation causes an energy change that restores the
system to its normal state, but under certain conditions the system is unstable. The
system can go from one energy equilibrium state to another. Some of these energy
states are stable within certain confines, but the operating zone can be induced to
move from one local minimum to another. This is illustrated diagrammatically in Fig.
A9.14, where the local equilibrium is illustrated as moving from B to A.
Consider an example. A perturbation in kinetic energy available to the yarn in the
balloon usually leads to a change in radius of the centroid and the yarn tension
changes in sympathy. If there is a change in mode, there is also a change in height
between nodes. The yarn is no longer roughly parabolic but assumes a sinuous shape.
There are a variety of balloon shapes in which the balloon contains the same amount
of yarn but has a different position of the centroid. If the perturbation acts perversely,
the effects permeate the system. If, for example, there is a decrease in yarn tension
at the surface of the bobbin due to a change in yarn shape similar to that discussed,
there is a reduction in energy available. If the reaction force between the ring and
traveler increases because of the reduced tensions, more energy will be dissipated
due to friction. This leaves even less available to the kinetic energy of the main
balloon, which then causes the balloon to deflate. The reduction in balloon diameter
further reduces the tensions and the system is seen to be unstable.
Kinetic energy
A9.6.2 Vector analysis
In standard ring spinning, where the machine designer does not intend collapse, the
event causes difficulties. For example, when the balloon is long, a portion of it
sometimes temporarily collapses on the top of the bobbin. When conditions verge on
instability, the balloon collapses periodically as the wind-on point approaches the top
of the bobbin. Often the result is that there is an end-break (which has economic
repercussions) or there are periods of increased yarn hairiness while the balloon
remains collapsed (which has quality repercussions). With coarse counts, the balloon
may reach a size that is over double that of the ring diameter before collapse occurs.
With a fine yarn, the balloon might collapse at a diameter roughly equal to the ring
size.
Let a yarn consist of a series of small elements and, for the present purpose,
consider the middle element in a chain of three. Figure A9.15 shows the forces acting
B
A
Balloon parameter (ωTrw )
Fig. A9.14
Various energy states
Advanced topics VII: Yarn balloon mechanics
449
Length of
element = δs
F + δ F1
F + δF 1
Fe 1
δγ
Fe 2
F δs / R
O
F – δ F2
R
δγ
Fe
R
Fe3
F – δ F2
Fe
Force diagram
Space diagram
δF and δγ exaggerated for clarity
(a) Case 1
F + (–δF1)
F + (–δF1)
F e2
δγ
O
F δs / R1
F – (–δF2)
δγ
Fe
Fe3
Fe
F δs / R
Force diagram
R
R
F – (–δF2)
Fe 1
Length of
element = δs
Space diagram
δF and δγ exaggerated for clarity
(b) Case 2 for negative δF
Fig. A9.15
Forces on an element of yarn
on the subject element of yarn (shown shaded in gray). Before discussing the meanings
of these, let the symbols be explained. For example, in diagram (a), a force of F + δF
is applied by the element immediately above it, and another force of F – δF is applied
by the element immediately below. This is regarded as a positive tension gradient.
The radius of curvature of the element is R and the forces all lie in the plane of
curvature, but they have been rotated about the element to make Fe horizontal in the
diagram.3 The vectors do not necessarily lie in the plane of the ring or in one including
the spindle axis. The external forces acting on the element are almost horizontal.
External forces are the centrifugal force acting on the element and the airdrag forces.
The latter is true only if any secondary airflow produces negligible airdrag on the
element. Thus by rotating the plane about a vertical axis to make Fe parallel to the
resultant of the external forces, the element is brought into its correct attitude in the
balloon.
In Fig. A9.15(b) the tension gradient is negative (the value of δF is negative) and
the attitude of the element has changed in consequence. Resolve the principal
3 Constraints are (a) Fe1 = Fe2 + Fe3 and (b) the sum of the moments about any point on the
element has to be zero (which implies that Fe2 ≠ Fe3). The moment arm about which Fe acts is
not δs/2, unless Fe and Fe1 are coincident.
450
Appendix 9
components shown in gray. In the right-hand diagrams in the direction O–Fe1, the left
facing components Fe2 and Fe3 (shown in black and facing leftwards) are unequal.
Consequently, there is a moment tending to tilt the element, which should be balanced
by the moment generated by the application of Fe. Thus, the tension gradient along
the yarn within the balloon affects the attitude of the element with respect to its
neighbors. Clearly, Fe is greatly influenced by any change in the radius of curvature.
The behavior of the balloon is influenced heavily by changes in radius of curvature
and tension gradient.
Equilibrium of the balloon occurs only when the outward forces balance the inward
ones. The outward forces are a combination of the centrifugal and drag force acting
on the yarn element. Drag forces are mostly tangential to the locus and have little
direct effect on this balance when the balloon is fairly upright. However, collapse is
initiated in the region just above the ring where airdrag causes the yarn in the balloon
to incline almost to its maximum extent. A rough approximation in that zone is to
treat the plane of curvature as the same as that of the ring, which implies that the
radius of curvature is smaller than elsewhere in the balloon. The consequence is an
increased tendency to reduce the radius of the locus in the lower regions near the
traveler.
A9.6.3 Collapsed balloon spinning
It is fairly obvious that if the radius of the yarn balloon could be reduced as a
practical proposition, the yarn tensions could also be reduced. In long-staple spinning
this is an option, but in short-staple work it is not. Generally, collapsed balloon
spinning is used for heavy, long-staple yarns that are capable of withstanding high
tensions. The friction tends to make the yarns hairy. The winding tension in such
cases is partly determined by the friction between the sliding yarn and the machine
surfaces. It is also partly determined by the end conditions, which are determined by
the traveler weight and other parameters already discussed. Lubrication of sliding
surfaces is also a factor.
If the tensions are properly adjusted, it is possible to make the balloon collapse, or
run at a reduced size, as shown in Fig. A9.16. Although the centrifugal component is
much reduced by this, there is now a significant frictional drag as the yarn passes
over the surface of the spindle or crown. The frictional drag may be calculated
approximately from Amonton’s Law using Equation (A9.13); this implies that the
ratio of tensions is a function of the angle of wrap and the coefficient of friction. For
the system to pay off, the increase in yarn tension due to friction by the above
mechanisms must be less than the increase caused by allowing the balloon to inflate.
Usually, a crown is mounted on anti-friction bearings on top of the spindle to reduce
the frictional forces. However, the yarn still has to slide over the crown in a direction
along the length of the yarn.
A9.6.4 Unwinding
The ring bobbins provide only temporary storage and the yarn has to be unwound
from them in the so-called winding process as was described in Chapter 9. Winding
machines usually pull yarn over-end from a stationary package. The package from
which the yarn can be removed might be a ring bobbin, cone, or cheese, although the
most common is the ring bobbin. The yarn is caused to balloon by the motion of the
Advanced topics VII: Yarn balloon mechanics
451
From
drafting
system
Collapsed balloon
Fig. A9.16
Reduced balloon
Balloon collapse
wind-off point on the surface of the package. The take-off speed and the radius at
which the departing element of yarn is removed from the bobbin determine the
rotational speed of the balloon. Of necessity, the rotational speed is variable and the
structure of the package causes the take-off point to oscillate rapidly. The balloon
changes height, diameter, and shape as a result, and a chaotic balloon is created.
There is usually no ring and traveler to help smooth out the fluctuations. Any
instantaneous view of the yarn in the balloon shows a multi-noded sinuous shape
rotating about the package axis.
The presence of ballooning forces is important because they hold the yarn clear of
the surface of the package. This avoids the removal of neighboring coils of yarns that
would result in tangles being formed. It also reduces the amount of hairiness created
by the over-end unwinding process.
A9.7
Balloons in two-for-one twisting
A9.7.1 Tension control by the use of cylindrical surfaces
Two-for-one systems involve the high speed removal of yarn from large diameter
packages stored inside the balloon. Consequently, the shape of the balloon has to be
controlled to prevent the yarn just removed from rubbing the surfaces of the package(s).
The tensions have also to be controlled because of the high speeds and diameters
involved. Often there are two coaxial balloons involved and these have to be kept
separate. For these reasons the balloons are frequently contained within cylindrical
cans which act rather like balloon control rings.
A9.7.2 Tension control by friction devices
An absence of any frictional type of control leads to balloon instability and it is
normal to use a spring-operated tensioner or a governor operated by centrifugal
452
Appendix 9
forces to help control the tension. Changes in yarn length within the balloon are
accommodated by a disk designed to dynamically store limited amounts of yarn. An
extreme and undesirable case is that of the chaotic balloon just described in Section
A9.6.4, which lacks any such a control.
References
1.
2.
3.
4.
5.
6.
7.
8.
Shintaku, S. Oda, J and Yamazaki, H. Airflow around the Rotating Pirn (or Cop) and Power
Loss, J Text Mach Soc Japan, 43(1), T1–T9, 1990.
Lord, P R, Rust, J P and Fenercioglu, F. Balloon Irregularities in Ring Spinning, J Text Inst,
1997.
McAdams, W H. Heat Transmission, p 99, McGraw Hill, 1942.
Batra, S K, Ghosh, T K and Zeidman, M I. An Integrated Approach to Dynamic Analysis of
the Ring Spinning Process Part I, Text Res J, 59, 6, 309–17, 1989.
Batra, S K, Ghosh, T K and Zeidman, M I. An Integrated Approach to Dynamic Analysis of
the Ring Spinning Process Part II, Text Res J, 59, 7, 416–24, 1989.
Lisini, G G, Toni, P, Quilghini, D and Di Giogi Campedelli V L. A Comparison of Stationary
and Non-stationary Mathematical Models for the Ring-spinning Process, J Text Inst, 83, 4, p
550, 1992.
Grishin, P F. Fundamentals of Spinning Ring Development, Whitin Review, 25, 3, p 34, 1968.
Frazer, W B. Ring Spinning, Text Horiz, Benjamin Dent & Co. Ltd, pTH 37, 1996.
Appendix 10
Advanced topics VIII: Topics in rotor
spinning
A10.1
Brief history of open-end spinning
The idea behind open-end (OE) spinning is almost as old as history itself. Farmers
twisted straw into binders for stooks of corn and wheat by continuously adding new
straw to the end of the binder and twisting it into the existing structure to make it ever
longer. The industrial revolution saw some clumsy attempts at a mechanical solution,
but it was not until the twentieth century that elegant solutions began to appear [1].
Derivatives of two of the systems then envisioned have become established, namely
air-jet and rotor spinning.
Early patents by Götzfried [2] disclosed the idea of using an air vortex to assemble
fibers and twist them into yarn. Lord [3,4] worked on such vortex systems but fiber
losses and yarn structure were unacceptable; the method was then commercially
unattractive, despite the simplicity of the device. A design where the vortex was
confined to the fiber assembly was offered for sale but did not achieve significant
market penetration. It was left to Nakahara [5], Morihashi [6], and others to develop
a system that used air-jets to twist but allowed the fiber assembly to be controlled by
other processes. Although the idea started out as a sort of OE spinning, the successful
system lost the essence of OE spinning because there was no longer an open end,
merely a very ingenious way of manipulating twist and yarn structure. This was
described in Chapter 10 and it is merely a matter of peripheral interest in this context.
The origins of rotor spinning were in the work of Berthelson [7] in 1937 and
Meimberg [8]. In the early 1960s, VUB [9] in Czechoslovakia, the Shirley Institute
[10] and UMIST in Manchester, UK, SRRL in New Orleans, USA, and perhaps
others, were experimenting with rotor spinning. VUB produced a working prototype
designated the KS200, which was the predecessor to the BD200, which was offered
for sale in 1966 at $200 per spindle. Eventually, the Czech BD200 gained a good
market share.
In the very early days, there seemed to be a limit of about 20 000 r/min in rotor
speeds, because of bearing design. Also, calculations of that time suggested that a
prudent speed limit might be about 25 000 r/min for a 3 inch (76 mm) diameter
aluminum rotor with air pumping holes. (The pumping holes introduce stress
454
Appendix 10
concentrations that reduce the strength of the structure.) A number of experimenters
found that rotors deformed or even burst when oversped. Somewhat later,
Landwehrkampf [11], who was concerned with large rotors for long-staple rotor
spinning, opined that plain 120 mm (4.75 inch) diameter rotors could run at 25 000
r/min. In a study by Wunsch [12] and Kerr [13], the energy consumption of a plain
disk was found to be proportional to D3.8 ω2.5 and this implies that large rotors
running at high speed will get very hot. Landwehrkampf published some curves that
showed a 5.46 inch (138 mm) rotor running at 20 000 r/min required nearly 300 watts
and approximately 1.5 inch (38 mm) rotors required about 120 watts to run at 80 000
r/min. However, by the time losses are included, a frame of 100 large rotors running
at 20 000 r/min seemed to require well over 30 kW. The data of Landwehrkampf did
not fit those of Wunsch and Kerr but that was not surprising because of the differences
in shapes of the rotating member. For various reasons, the long-staple rotor spinner
did not succeed commercially and it was the short-staple version that made a remarkable
impact on the industry. From a consideration of the foregoing, it was estimated in the
1970s that a machine of 300 rotors of 1.5 inch (38 mm) diameter running at 100 000
r/min would require some 60 kW. This is a very large power demand and it was clear
that the rotor size had to be reduced. The size of 38 mm had been picked on the basis
that the diameter of the rotor ought not to be less than the fiber length for quality
reasons. This idea was proved wrong as it turned out. In 1997, 28 mm rotors could
be run at 130 000 r/min (the power consumption is not known to the author) and
commercial speeds ranged between 85 000 and 110 000 r/min. The reduction in rotor
size has continued over two decades and it certainly has been connected with the
rising power demand at the ever higher speeds. In the modern rotor machine, the
temperature of air leaving the rotor is very high. How much further these trends can
go is another matter.
The early experiments at UMIST [14] indicated that rotor spun yarns were weak
in comparison to ring yarns. One reason was the incidence of bridging fibers that
caused an enlarged population of hooked fibers in the yarn. For this reason it was
concluded (wrongly) that, to make the yarn attractive to spinners, the circumference
of the rotor had to be many times the fiber length, in order to reduce the proportion
of bridging fibers. The elegance of the tapered rotor sliding wall, which conserves
space for the assembling of fibers, was not appreciated then.
The history of ring spinning shows the difficulties of getting a new process accepted,
and one can find old articles opining that ring spinning would never replace the mule.
Similarly, rotor spinning took time to become established and yarn weakness was one
of the reasons for the reluctance. The market then began to accept the yarn for what
it was, but it still took many years before it was fully accepted [15,16]. The relatively
low cost of operating rotor spinning has always been one of its main attractions
although the capital cost per rotor was initially four or five times that of a comparable
ring spinning machine position. Consequently, the rotor had to be run faster to reduce
the capital cost/lb of yarn produced to competitive levels. It is the history of this
pursuit of speed that is so fascinating. Small rotors, new alloys, protective treatments
to withstand wear, and new drive systems all made their contributions. The driving
force behind this was to reduce the capital cost/lb of yarn by increasing rotor speed.
It was necessary to increase productivity faster than capital cost in order to achieve
this. Many of the doubts and reservations of the time are well expressed in a review
of rotor spinning made in 1978 [17].
Advanced topics VIII: Topics in rotor spinning
A10.2
455
Yarn evenness
A10.2.1 Number of doublings inside the rotor
Fibers are laid into the vee-shaped collecting surface inside the rotor, and enter as a
thin stream of fibers. It takes many layers of fiber to make up sufficient linear
density; in other words, there are many doublings. These doublings tend to even out
any short-term irregularities in the yarn and OE yarns tend to be surprisingly even.
Also, there are no errors carried forward from a roving frame, and many errors
created by the combing roll drafting system are smoothed. However, longer-term
errors arising from the sliver still remain, and these are usually neither worse nor
better than with ring yarn.
Referring to Fig. A10.1, let n = linear density of the fiber, V = velocity, M = mass
flow, m = number of fibers in the cross-section, and the subscripts f and y refer to
fiber and yarn, respectively. Also let ω = rotational speed in rad/sec, r = radius of
collecting surface of rotor, and τ = twist/unit length of yarn.
Mass flow/unit time at input = Mf = mf n Vf
Mass flow/unit time at output = My – my n Vy
But
M f = My ,
from which:
my/mf = Vf/Vy.
If Vf = ωr and Vy = ω/2πτ, then
my/mf = 2πrτ
[A10.1]
= number of internal doublings in the process
Thus, for a 30 tpi yarn running in a 1.5 inch diameter rotor, there are approximately
140 doublings in the rotor groove. This, then, is why the short-term unevenness is so
good in comparison to ring yarns.
my Fibers in
cross-section of yarn
mf Fibers in
cross-section of flow
My
Vy
ω
r
Vr
Fig. A10.1
Vf
Mf
Conservation of mass flow in the rotor
456
Appendix 10
A10.2.2 Short-term blend evenness
Multiple doublings inside the rotor improve the short-term evenness and the intimacy
of the blend [18]. In theory, the per unit CV of linear density of the yarn should be
√(my/mf) for lengths up to (my/mf) × rotor circumference. Figure A10.2 is from the
work of Deshpande [19], who blended dyed viscose rayon with polyester fibers at a
single passage of drawing, before spinning the blend on an OE machine. The machine
used is now obsolete but the work shows that the multiple layering inside the rotor
ensures a good blend. However, careful examination of Fig. A10.2 shows several
spots where there are concentrations of similar sorts of fiber, and the homogeneity is
not as perfect as might be hoped. Good dispersion of the components requires that the
slivers be properly prepared and that the combing rolls in the OE machine be maintained
and operated correctly.
A10.2.3 CV of linear density
With staple yarns, the number of fibers in the cross-section can vary considerably. If
we assume that there is a Poisson distribution in this number, mav is the average
number of fibers in the cross-section, m is the actual number, and s is the standard
deviation then:
[A10.2]
CV = s/mav
The standard deviation for this type of distribution is estimated to be a function of m;
hence, if the value of CV is not too large:
CV = 100/√m %
[A10.3]
But m is related to the linear densities of yarn and fiber. Thus it will be realized that
the CVs of blend, strength, and count vary with linear density. A 36s cotton yarn
made from 4.5 micronaire fibers only has about 93 fibers in the cross-section and we
Fig. A10.2
Cross-section of a blended rotor yarn
Advanced topics VIII: Topics in rotor spinning
457
would expect 10.4% CV due to randomness of the fibers. If the yarn had been made
of 2.5 micronaire cotton, the number of fibers would have increased to about 167 and
the CV would reduce to 7.7%.
Components due to organized errors arising from malsetting of the machines and
variations in fiber properties should be added vectorially to these figures. The effect
of doubling is to bring the actual CVs closer to the minimum values. Since the rotor
doubles over a length equivalent to the rotor circumference, one can expect that
errors shorter than, say, 6 inches (150 mm) will be sufficiently doubled and the shortterm evenness should be improved. However, the actual short-term error is still
significantly above the theoretical values predicted by Equation (A10.3) (but is normally
better than with ring yarns). Some reasons for this are discussed in the next section.
The long-term errors are little affected by the doubling in the rotor and are dependent
on the doubling at the drawframe and other preparatory machines. If a single passage
of drawing is used with eight slivers in the creel, there might be only eight doublings
there. This is much less than is found within the rotor. The point being made is that
preparation has a larger relative impact on long-term yarn evenness with OE yarn as
compared to ring yarn. It will be recalled that poor preparation can induce high error
production in ring spinning.
A10.3
Toothed drafting
A10.3.1 Combing roll clothing
Combing rolls pull fibers from the beard of a sliver that is continuously fed by a feed
roll and plate system. The combing roll usually rotates between 500 and 9000 r/min,
it is clothed with either saw-teeth or needles, and its function is to detach fibers from
the advancing fiber beard. If the sliver is not well prepared, fiber breakage can ensue
because the entanglement of fibers in a clump increases the withdrawal force per
fiber and more are caused to break than is desirable. Some measurements were made
with rayon fibers on an old OE spinning machine. Undyed rayon fibers can be made
almost invisible in a bath of liquid methyl salicytate so that a dyed tracer fiber within
the structure of the yarn can be seen among the surrounding fibers. Dyed tracer fibers
were placed carefully in the sliver entering the OE machine and the yarn produced
was studied under a microscope. In the yarn, the fiber extent (the distance between
the extremities of a folded fiber embedded in a yarn) was greatly reduced as the
fibers took up a variety of hooked and looped shapes. Sometimes the original fiber
was found to exist in two or more pieces and often only a shortened piece of tracer
fiber would be found. When the fiber placed on the sliver was greatly crimped or
relaxed into a very convoluted shape, the fiber almost invariably broke. The condition
of the combing roll wire, its speed and its shape, all affected breakage rates; it also
affected the ejection rate for trash.
Siersch [20] showed that helicoidally arranged teeth on the combing roll split fiber
tufts into roughly parallel fibers separated by contiguous teeth (Fig. A10.3). This
beneficial separation was accompanied by cyclic fluctuations in fiber flux (number
of fibers per unit area of flow) and yarn tension, which were related to the pitch of
the tooth helix. CVs of the fiber flux in his experiments varied between 8.9% and
9.6%. This, then, accounts for one of the reasons why the short-term CV is greater
than the theoretical value. The larger the number of tooth helices, the smaller was the
variation, and the higher were the yarn strength and breaking elongation. Too fine a
458
Appendix 10
Fiber
Motion of teeth
Penetration of combing roll teeth into a fiber beard
Fig. A10.3
tooth pitch (< 2 mm) created increased nep production with cotton fibers and a
deterioration in yarn CV (Fig. A10.4). If the front angle of saw-tooth clothing (Fig.
A10.4) was increased above about 20°, the drafting force increased and so did fiber
damage. Various investigators have shown that combing roll damage produces yarn
irregularity. The most usual damage is to the teeth; sometimes careless handling
causes this, sometimes it is caused by large particles in the feed sliver, and sometimes
by fiber jamming. The latter can be caused if a loop of sliver is lifted from the can and
a double, or triple, thickness of sliver is ingested by the feed roll. A common time for
this to happen is when a can is being emptied of the last length of sliver. However, it
can happen when a sliver piecing has just been performed, or if a can has been
damaged.
The life of the combing roll clothing is finite and the use of dusty fiber, or of fiber
with abrasive fiber finish, increases the wear rate on the teeth. Consequently, not only
are the metal surfaces hardened, but they are also surface treated to improve their
wear resistance. Like card wire, the body of the tooth has to be tough to prevent
brittleness; thus, despite the hardness of the cutting edge, the teeth can be bent. Bent
teeth result in a loss of evenness in the yarn.
A10.3.2 Combing roll bearings
Combing roll bearings become damaged in service. Slippage in the tape drive can
cause the bearings to become overheated, which causes the grease to fail. Typically,
the grease hardens and blocks further lubricant from reaching the ball track.
1.5
17
1.0
16
CV yarn
0.5
Nep/sq inch
Yarn CV (%)
α
Nep/sq in
15
–20 –10
Fig. A10.4
0 10 20
Tooth angle (α)
30
0
40
Effect of comber roll tooth angle on yarn performance
Advanced topics VIII: Topics in rotor spinning
459
Shock or overloading can cause the balls to indent the ball track. The race becomes
noisy and consumes more power, which, in turn, leads to lubricant failure. Tests [21]
using accelerometers to measure the vibrational accelerations at the combing roll
bearing housings showed unusually high values for worn units. The use of an encoder
driven by the combing roll enabled the vibration pattern to be resolved, and a sample
is shown in Fig. A10.5. Cutting the bearing housings open revealed damage to the
ball tracks.
A10.4
Fiber assembly – the formation of wrapper fibers
Once per revolution, the laying of the fibers on the collecting surface and the peeling
of the yarn from the collecting surface interferes. Fibers laid at these times are called
bridging fibers and, during removal, portions of these fibers become bent back and
wrapped around the body of the yarn.
Consider Fig. A10.6(a). A fiber is shown sliding on the inside of the conical
portion of the rotor. One end is already trapped in the yarn leaving the rotor groove.
The peeling point is where the yarn leaves the rotor groove. As yarn is pulled from
the rotor, this peeling point should move in the same direction as the rotor. The dotted
line represents a sliding path of a fiber in the recent past. In Diagram (b), events
occurring very shortly after the first are portrayed. A small amount of yarn has been
removed, carrying the entrapped fiber with it; meanwhile, the trailing end of the fiber
has slid nearer the rotor groove. Eventually, the trailing end must be folded back on
the core of the yarn, as depicted in notional form in Diagram (c). However, the yarn
rotates about its axis because of the false twist, and this causes the folded back fiber
to become wrapped around the core of the yarn as indicated in Diagram (d). Variation
in inclination of the fibers within the yarn is typical of the structure. Portions of the
bridging fiber are wrapped around the outer surface of the yarn and carry very little
load when the yarn is in the free state (i.e. not assembled into fabric). The remaining
portions of the bridging fibers are buried in the yarn structure and, although they
carry some load, they behave like short fibers. Consequently, the yarn is weaker than
Pulses due to the indentations
in the tracks and balls
Ball damage
Outer track
Shaft
Inner track
Periodic track
damage
Fig. A10.5
Polar diagram of acceleration at
the bearing housing
Combing roll bearing damage
460
Appendix 10
Fiber
Fiber
Fiber pulled from
the rotor by the
departing yarn
Original fiber
path
Yarn
Yarn
Fibers sliding
inside the rotor
Tail slides
(a)
(b)
Hooked fiber
Yarn
Twisted hooked fiber
(c)
(d)
Fig. A10.6
Bridging fibers
ring yarn. Wrapper fibers increase the pressure on the enclosed fibers and this gives
some local resistance to failure, although it produces an unwanted waisting in the yarn.
The chance of a bridging fiber depends on the rotor diameter and the projected
length of the sliding fiber approaching the rotor groove. The term ‘projected fiber
length’ must be explained. The fiber does not approach the rotor groove with its
length parallel to a tangent of the rotor groove; rather, it approaches obliquely.
Furthermore, the fiber may not be straight but might be convoluted in some way.
Thus, if viewed perpendicular to the direction of slide, the distance between the
extremities of the fiber is less than the real length. This distance between the extremities
is referred to here as the projected length. To repeat, it is always less than the actual
fiber length. Consider an example where the circumference of the rotor is 3 inches
and the projected fiber length is 1 inch. In such a case, two out of every three fibers
will be assembled inside the rotor groove without intersecting the path of the outgoing
yarn. One in three will intersect the outgoing yarn and might be entrapped by it. The
first, and larger, category of fibers becomes the core of the yarn and the second
category become wrappers. The structures of the yarn are discussed in Appendix 5.
Figure A10.7 shows some micrographs in which the core has been shaded to highlight
the wrapper fibers. The wrappers are shown as dark fibers.
A10.5
Twist distribution
A10.5.1 False twist control by use of a rotating navel
Causing the navel to rotate can change the false twist created at the navel. Lünenschloss
[22] showed that using a rotating navel increased the minimum TM at which one
could spin (Fig. A10.8). He also showed that a soft yarn with a low twist could be
spun. This was attractive not only because of the hand of the yarn but also because
a low twist multiple gives a potential productivity increase. However, the rotating
navel was an extra complication and it has not proved acceptable in practice.
Advanced topics VIII: Topics in rotor spinning
461
(a)
(b)
(c)
Fig. A10.7
Rotating
Minimum twist multiple
6
Fixed
5
4
3
Rotor diameter = 50 mm
Linear density = 50 tex
2
20 000
Fig. A10.8
Wrapper fibers
40 000
60 000
Rotor speed (r/min)
80 000
Effects of fixed and rotating navels in rotor spinning
A10.5.2 False twist distribution in the rotor vee
The geometry of the navel affects the performance. As explained in the main text, the
yarn rolls on the navel and creates false twist in the yarn inside the rotor. It should be
noted that the twist of the yarn arm inside the rotor can be significantly higher than
that in the emerging yarn. The torque of the yarn in the rotor vee is relieved at or near
the peeling point of the yarn. A length of incipient yarn lying in the rotor vee adjacent
to the peeling point has a varying level of twist, as indicated in Fig. A10.9. The shape
of the rotor groove affects this twist propagation and a sharp vee tends to restrict the
propagation more than a rounded one. An approximate distribution of twist in the
incipient yarn lying in the rotor groove is:
T ≈ To ekµθ
[A10.4]
where µ = coefficient of friction, θ = angle subtended by the incipient yarn measured
from the peeling point, k = ω2r2n/2 sin α and 2α = angle of the rotor groove. Thus,
the groove angle has a strong effect on performance. If α is too small, the yarn jams
in the groove.
The distribution of forces acting on the yarn lying in the rotor groove is shown in
Appendix 10
Total twist in
rotating yarn arm
Yarn twist
462
Twist in emerging yarn
Length along the yarn inside the rotor
Twist distribution inside a rotor
Fig. A10.9
Fig. A10.10. A rounded groove gives a different distribution of forces, reduces the
total lateral force acting on the yarn from F to F′, and modifies the coefficient k.
Wear can sometimes convert one shape of vee to another; this causes changes in yarn
characteristics. In the 1970s, before adequate wear protection treatments had evolved,
cases were known where the wear was sufficient to penetrate to the outside of the
rotor. Build-up of dust and trash inside the rotor also changes the shape of the vee and
affects the yarn characteristics.
A10.5.3 Twist surges
At very high speeds, difficulties begin to appear in retaining the twist in the rotating
yarn arm. False twist can surge forward through the navel, leaving a transient depletion
in twist inside the rotor, which causes an end-break near the navel inside the rotor.
Twist traps in the yarn withdrawal tube become a necessity at high speeds, especially
when spinning polyester or similar fiber. Many twist trap designs can be recognized
by the cranked doffer tube, which causes the yarn to leave at an angle to the rotor
center line. This is an important device in controlling the twist surges. The effect of
such surges can be recognized by the presence of portions of yarn inside the rotor
after an end-break. Without significant surges, there is only fiber and dust present
because the failure under non-surging conditions occurs where the yarn is peeled
δF = ω2r2nδθ
δFt = δF/2 sin α
δF t
δF t
α
α
α α
µ δF t
µδFt
Rotor vee
δ Ft ′
δF
δFt = δF/2 Sin α
Force diagram
(a)
δ Ft ′
µΣ δ Ft
µ Σ δ Ft
Fouled or damaged rotor
(b)
Fig. A10.10
Forces involved in rotor wear
Advanced topics VIII: Topics in rotor spinning
463
from the rotor groove. The angle of the cranked yarn withdrawal tube is important.
Normally it is cranked at about 45°. The smaller the angle, the lower the spinning
tension, but a torque-stop effect can be produced if the angle is increased. There are
also other designs of twist trap to fulfill a similar function.
The design and condition of the navel, as well as the character of the fiber and the
yarn count, play important parts in determining the nature of the yarn. The navel also
plays a part in determining the effectiveness of the operation. Yarn tension creates
forces between the orbiting yarn and the stationary navel. Normal forces between
yarn and metal create friction; the frictional forces act tangentially on the yarn and
produce torque. The mean tangential force is a function of the normal forces referred
to and the coefficients of friction between the surfaces. The normal force is proportional
to ω 2 rr2 n , where ω is the rotational speed, rr is the radius of the rotor, and n is the
linear density of the yarn. The false twist depends on the fiber finish, the type of
navel surface, as well as on rotor speed and size. An increase in coefficient of friction
decreases yarn strength but improves end-breakage rates. The flare radius connecting
the bore to the front surface of the navel plays a significant part in determining both
the false twist and the properties of the output yarn [13,14,15]. With a large flare
radius, it is possible to increase the false twist at the expense of the winding tension.
Lünenschloss et al. [23,24,25] wrote that the output yarn tension can exceed the yarn
strength when using a rotor diameter of 60 mm (2.36 inches) at rotor speeds above
70 000 r/min. This gives some idea of the problems at high rotor speeds.
A10.6
Conclusion
In high speed rotor spinning, attention to the design and condition of the combing roll
clothing is needed to preserve yarn quality. Attention to the design and state of the
rotors is important, not only to obtain high quality, but also to minimize the costs
associated with unnecessary end-breaks, and the consumption of power. Cleanliness
and maintenance are of great importance.
References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
Lord, P R. The Prospects of Break-spinning, 11th Canadian Textile Seminar, Kingston, 1971.
Götzfried, K. USP 2 926 483, 1955, German Pat 1115163, 1961 and BP 880239, 1961.
Lord, P R. Spinning from the Card using an Air Vortex, Textile Inst Ind, Jan 1967.
Lord, P R. Application of the Air Vortex Method of Spinning, Czech Sci and Tecn Symp,
Prague, Czechoslovakia, 1967.
Nakahara, T. USP 4 142 354, Mar 1979.
Morihashi, T. USP 4 183 202 Jan 1980.
Berthelson, S E. BP 477 259, 1937.
Meimberg, J. BP 695 136, 1953 and several others up to 1965.
Rohlena, V. Open-end Spinning, Elsevier , Oxford, UK, 1975.
Catling, H. The Comparative Merits of the Principal Break-spinning Systems, Textile Inst
Ann Conf, 1968.
Landwehrkampf, H. General Purpose or Special Machine for Open-end spinning, Melliand
Textilberichter, 4, 11, 1976.
Wunsch, H. Frictional Torque in Small Ball-bearings at High Speeds, NEL Report No 25
1962.
Kerr, J. Private communication, NPL.
464
Appendix 10
14.
15.
Lord, P R. Developing Rotor Break Spinning, Text Ind, Feb 1970.
Lord, P R. Commercial Developments in Open-end Spinning, Textile Inst Ann Conf, Harrogate,
UK, 1976.
Coll-Tortosa, L. Neue Aspekte der Technologischen Spinngrenzen des OE-Rotorspinnens,
Vorträge anlässlich der 3. gemeinsamen Tagung der Aachener Textilforschunginstitute, 1977.
Hunter, L. The Production and Properties of Staple-fibre Yarns made by Recently Developed
Techniques, Text Prog, 10, 1/2, 1978.
Lord, P R. Yarn Evenness in Open-end Spinning, Text Res J, 512–15, 1974.
Deshpande, S V. Open-end Spinning as a Means of Blending, MS Thesis, NC State Univ, NC,
USA, 1972.
Siersch E. Ein Beitrag zum Mechanismus der Fasertrennung und des Fasertransporte beim
OE-Rotorspinnen, Fortschrift-Berichte der VDI Zeitschriften, 3, 56, 1980.
TE402 Project, NC State Univ, Raleigh, NC, USA, 1995.
Lunenschloss, J. Private Communication.
Lunenschloss, J. Seminar Series at NC State Univ, Research into, and Design of, Rotor
Spinning Machines, NC, USA, 1975.
Lunenschloss, J, Coll-Tortosa, L and Phoa, T T. Die Einfluss der Faserlange und der Faserlangenverteilung auf die Eigneschaften Rotorgesponnener OE-Baumwollgarne, Textil-Praxis, Sept
1974.
Lunenschloss, J, Coll-Tortosa, L and Phoa, T T. Die Untersuchung der Faserstromung im
Faseleitkanal einer OE-Rotorspinnmaschine, Chemiefasern/Textil-Industrie 24/76, 355–485,
1974.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
Index
Abbreviations
fil = filament
HOK = normalized productivity
ls = long staple
m.m. = man made
m/c = machine
r.h. = relative humidity
spg = spinning
ss = short staple
2-for-1
2-for-1
2-for-1
2-for-1
2-for-1
2-for-1
2-for-1
2-for-1
2-for-1
2-for-1
2-for-1
2-for-1
2-for-1
2-for-1
2-for-1
2-for-1
2-for-1
co-axial balloons, 451
twisting, 220
twisting, armature, 256
twisting, balloon rail, 255
twisting, cylindrical pot, 255
twisting, fly & dust, 257
twisting, high speed, 66
twisting, package rotation, 65
twisting, piecing, 66
twisting, prep costs, 305
twisting, rotor-spun yarn, 255
twisting, space available, 255
twisting, storage disk, 257, 452
twisting, tension control, 451
twisting, yarn tension, 256, 257
twisting, yarn unwinding, 257
twisting, yarn waxing, 257
abrasion, 128
abrasive additives, 48
abrasive fibers, 417
acceleration zone, 70
access to capital, 313
accumulations of grease, 128
acetate fibers, 38
acid-cracking, 212
acrylic fibers, 21
acrylonitrile, 21
actual draft ratio, 321
added variance in drafting, 353
additive tensioner, 69
adequate quality of product, 350
advanced fiber info system (AFIS), 368
aerobic bacterial action, 212
AFIS, 285, 392
aggregations of fibers, 423
agricultural fiber production, 11
AGV, 184, 185
air conditioning, 148, 151, 283, 288, 289, 341,
346, 348
air ducting 128, 148
air permeability of sample, 353
air quality monitoring, 348
air vortex spg, 186
air wash, see air conditioning
air/steam mixtures, 343
air-drag, 438
air-jet spg m/c, 106, 107, 109, 262–267
air-jet spg ply yarn, 265
air-jet spg wide ribbon, 266
air-jet spg, fiber requirements, 264
air-jet spg, perforated surfaces, 262
air-jet texturing, 106, 266, 267
air-jet texturing, wetting process, 107
air-jet yarn, 89, 261, 267
air-jet yarn, polyester/cotton, 267
air-jet yarn, structure, 260, 267, 378
air-jet, yarn character, 107, 265–267, 378
air-jet/mechanical twister comb, 266
airstream study, 289
alternating self-ply, 68
alternating twist systems, 67
amino acid side chains, 31
amortization costs, 8
animal excretions, 208
anthrax, 32
anvil roller, 51
aperiodic nubs or loops, 109
aperture in cover plate, 262
aphid, 25
apparel market, 205
apron drafting, 79, 80, 174
aramid fibers, 22
artificial lighting, 349
466
Index
assignment specifications, 306
atmosphere, temperature control, 384
autoclave, 50, 90, 342, 343, 344
autodoffing, 175, 217, 305, 313
autoleveler, 86, 135, 136, 158, 159, 217
autoleveler, CPU, 136
autoleveler, echo, 135
autoleveler, temporary sliver storage, 136
automated guided vehicles, see AGV
automatic creeling, 175
automatic doffing, see autodoffing
automatic fiber transfers, 87
automatic handling, 9
automatic piecing, rotor spg, 310
automatic ratch setting, 420
automatic traveler changing, 313
automatic weighing system, 213
automation, 6, 184, 253, 303, 310, 313, 316
availability of capital, 5
availability of product, 301
average transit time, spindle-spindle, 308
axiflow machine, 123, 127
back-leakage in extruder barrel, 42
back-pressure of filter pack, 42
backwashing, 216
bacteria & viruses, 25
bad bale laydown, 404
balance, labor cost & capital, 7
balanced plied yarns, 60
bale blooming, 120, 121, 123
bale conditioning, 120, 216
bale cutter, 121, 122
bale decompression, 31
bale density, 31, 121
bale height, 121
bale income, seasonal, 391
bale laydown, 72, 121, 122, 131, 392, 394
bale laydown, withdrawal consistency, 391
bale lots, 119–121
bale management, 118, 120, 122, 129, 391,
392, 394
bale milling, see bale plucker
bale mixing, 397
bale plucker movement, 131, 167, 395, 397,
403, 404
bale plucker, 117, 121, 394
bale plucker, depth of cut, 131, 397
bale plucker, productivity, 131
bale preparation, 119
bale profile, 395
bale size, 13, 31
bale storage, mill, 392
bale ties, 31
bale wrapping, 31, 113
bale, assignment to category, 391
bale, compressed, 117, 120
bale, fiber tags, 122
bale, selection from warehouse, 119
bale-bale variation, 116
bale-plucker, cyclic removal of fiber, 397
bales issued, category limits, 392
bales, between-lot variations, 120, 391
bales, fiber attributes, 391, 395, 405
bales, fiber removal, 395
bales, moisture content, 120
bales, selection for laydown, 392, 393
bales, ss fibers, 14
bale-to-bale variation, fiber attributes, 131
baling fibers, 53
balloon collapse, 181, 299, 441, 448
balloon control ring, 178, 436, 441, 442
balloon geometry, 178, 427
balloon, 2-for-1 twisting, 451
balloon, common axis, 65
balloon, real, 436
balloon, yarn tension, 427
ballooning, length of streamers, 438
bargaining, supplier and customer, 276
barré, see fabric barré
barrel/screw clearance, 43
barrel/screw cooling seizure, 43
basis weight, 371
bast fiber spg, 231, 232
bast fiber spg, gilling, 232
bast fiber spg, hackling, 231
bast fiber spg, roughing, 231
bast fiber spg, shives, 232
bast fiber spg, spreading frames, 232
bast fiber tow, 231, 232
bast fiber wet-spg, 232
bast fiber, cuts (hanks), 232
bast fiber, dry-spg, 232
bast fiber, grist (count), 232
bast fiber, leas (hanks), 232
bast fiber, rove (roving), 232
bast fiber, sampling procedure, 279
bast fiber, spread sheet, 232
bast fiber, working, twist & untwist, 232
bast fibers in stem bark, 2, 35
bast fibers, history, 2
beat frequencies, 398
beaters, 123
beetle larvae, eggs, 32
belt transmission calc, 332–334
between-stream variance, 423
bi-component sheets, slit, 114
bicomponent yarns, 85, 110, 111
bimli (hibiscus cannabinus), 36
biochemical oxygen demand (BOD), 212
biological control, 32
blend component discrimination, 390
blend CV, 390
blend evenness, 164, 389, 392, 399, 402, 406
blend proportions, 117, 121, 400, 404
blend, fractionation, 390
blend, intimacy, 117, 129, 131, 155
blend, sudden changes, 400
Index
blending machine, 116, 117, 131, 132, 133
blending man-made fibers, 13
blending product streams, 84
blending, 13, 51, 116, 389
blending, bowl analogy, 389
blending, mixing before carding, 129
blending, slivers, 129
blends, wool/m.m.fiber, 14
blowroom waste disposal, 150
blowroom, 116–118, 150
blowroom, fiber losses, 150
bobbin (slubbing) transport, 230
bobbin (slubbing) wind structure, 230
bobbin feed to winders, 255
bobbin flow, 327
bobbin transfer systems, 155, 255
bobbin, 168, 169, 185, 230, 299, 444, 445
bobbin, random mixing, 133, 134
bobbin, unwinding, 185
bobbins, deformed, 185
boiler leaks, 344
boiling point, 342
bonds, molecular, 85
bottom roll fluting, 156
break draft, 72, 414
breast works, 213
brighteners, 48
Btu/hr, 342
bulk, see yarn, bulk
bulked continuous fil (BCF), 106
burr picker, see card, ls, burr beater
byssinosis, 154
cabling machine, 66
calculations, 317
calibration cottons, 355
camel hair, 31
can-changers, 155, 310
cans of sliver, transport, 310
cap spg, 218
capital & fixed costs, 309
capital cost per stream, 87
capital cost, 6, 50, 87, 291, 301, 303, 309,
310, 313, 348
capstan friction, 46, 75
capstan tensioner, 69
card autolevelers, see autoleveler
card clothing, 145
card licker-in waste, 153
card productivity, 324, 325
card sets, 15, 215, 219, 230
card waste, 149, 152
card, feed roll, 154
card, high inertia, 154
card, licker-in, 187
card, ls, blending, 214, 215, 224
card, ls, burr beaters, 210, 213
card, ls, carding actions, 213–215, 219, 224,
227, 228
467
card, ls, element sizes, 224
card, ls, element speeds, 227
card, ls, fettling, 227
card, ls, multiple worker/strippers, 215
card, ls, nep creation, 216, 228
card, ls, producion rates, 214, 224
card, ls, wire, 227
card, parallel streams, 398
card, roller-top, 15, 213–228
card, woolen, 224
carding m/c, ss, 118, 136, 137, 139, 141–147
carding prep, ss, 116
carding, ss, 136–139, 143, 144 –146
carding, ss, card wire see card clothing
carding, ss, clothing, 139, 141, 143, 145–147
carding, ss, control chart of nep, 147
carding, ss, cotton cleaning, 142–145
carding, ss, doffing, 136
carding, ss, dynamic testing of setting, 147
carding, ss, fiber condensation, 136, 142, 143
carding, ss, fine fibers, 147
carding, ss, grinding, 139, 141, 145
carding, ss, knife edge, see cleaning edge
carding, ss, nep, 136, 139, 141, 146, 147
carding, ss, performance deterioration, 141
carding, ss, re-clothing, 139, 148
carding, ss, rewiring see reclothing
carding, ss, screens, 144
carding, ss, segments, 139
carding, ss, separation of fibers, 136
carding, ss, settings, 146, 147
carding, ss, sliver can, 143
carding, ss, sliver, 142, 143
carding, ss, sticky cotton, 143
carding, ss, trailing fiber hooks, 143
carding, ss, trash separating devices, 144
carding, ss, trumpet or condenser, 142
carding, ss, wire see clothing
carding, ss, worn teeth, 139
carding, worsted, 213
cards in parallel, 119
cards, draft ratios, 73
cards, input fiber variation, 395
cards, ls, multiple-doffer split-web, 220
carpet backings, 113
carpet market, 205
carpet yarn manufacture, 219
caterpillar drafting, 216
ceilings, 349
cellulose tube, 22
cellulose xanthate, 4, 40
central signal processing, 350
centrifugal force, 169
ceramic rings, 182
cessation of flow, 43
change in state, 342
change the X-sectional area of strand, 70
changes in fiber content to variation in linear
density, 86
468
Index
channeling, 44, 84, 398
characteristic times, 395
charges that the market can bear, 309
chase, 177, 181
cheese, 114, 234, 237
chemical contamination, 349
chemical precipitation, 40
chemistry, 11
choice of fibers, 307
choice of solvent, 38
chokes, 76, 128, 174, 182, 224
Churka gin, 27
chute feed, 118, 122, 129
claim settlement costs, 392
clamping of hooked fibers, 353
classers’ length, 280
classes of economies, 5
classifying spinners, 286
clean air in mill, 348
cleaning machines, 117, 128
cleaning means, 127, 128
cleaning points, number, 127
cleaning, process, 125
clogging, spinneret, 38
clump-clump, fiber attributes, 390
coagulation, 4, 38, 40
coefficient of air-drag, 438
coefficients of variation (CV), 83
coercive torques, 384
co-extrusion, curl, 113
cohesion curve, 59
cohesion, shared fibers, 68
coil geometry, 386
coiler, 136, 149, 157
collapsed balloon spg, 219, 450
color matching, 216
color measurements, 404–406
color tests, knitted fabrics, 405
comb overloading, 163
combed wool sliver, 217
comber lap, ss, 159
comber noil, 162, 163, 286
comber production, 324
comber roll, OE clothing, 188
comber speed, 162
comber webs, 159
comber-roll, OE, 186–191, 425, 457–459
comber-roll, OE, abrasive fiber finish, 458
comber-roll, OE, life, 458
comber-roll, OE, loop of sliver fed, 458
comber-roll, OE, performance, 189, 191
comber-roll, OE, use of encoder, 459
combing elements, damage, 162
combing m/c, ss, 159, 161, 162, 163
combing process, 161, 161
combing, cost, 161, 301
combing, doubling, 163
combing, fiber dispersion, 397
combing, fiber orientation, 159
combing, maintenance, 162
combing, short fiber removal, 159
combing, sliver appearance, 159
comfort zone of operatives, 348
commercial nylon, 1940, 4
common principles, 56
compact spg, 261
competition between suctions, 262
competition, 87, 303
complaints, 277, 300
component balance, 162
component HOK & OHP, 304
composite yarns, 15, 260
compressed fibers buckle, 50
compression (nip) zone, 81
computers, local, 350
concentricity of drafting elements, 73
condenser, 81, 122
cone pulleys, 169
cone winding, 239, 240
cones or cheeses, damaged, 198
cones, 114, 234, 239
confined & non-confined systems, 67
coning oils, 95
conservation of computing power, 359
conservation of flow, 71, 119, 313, 320
continuous fabric inspection, 372
continuous fil, 11, 18
continuous heating systems, 50
continuous measurement, 356, 417
control & autoleveling, 135
control charts, 285, 286, 358
control limits, 351
control mechanisms, 71, 82, 342
control of flowing material, 71
control signals, 350
controlled climate, 344, 347, 349
controllers, 71, 133, 134
controlling number of categories, 392
conversion between count systems, 319
conversion cost, 301, 302, 311, 312
conversion of stems to sliver, 231
cored slivers, 81
corrosion in boilers, 344
cost & price, 301
cost & quality of product, 9
cost considerations, 209
cost data, 8
cost estimates, transitory, 309
cost minimization, 308
cost of fibers, 7
cost of labor, 308
cost of waste, 128, 150, 308
cost proportions, 301, 310
cost savings, end-break repairs, 306
cost, quality penalties, 306
cost, spg, 315
cost, what the market will bear, 392
cost, winding, 315
Index
costs & sales, 8
costs by percentage, 304
costs, historical, 303
costs, supplier, 276
cots, 75, 156, 174, 415
cotton & flax, 2
cotton & synthetic fibers, 3
cotton bale labeling, 279
cotton cleanliness, 189
cotton count X breaking load (CSP), 366
cotton fiber merges, 391
cotton fibers, cost, 5
cotton ginning, 27, 143
cotton growing, 25
cotton hank, 318
cotton module, 26, 279
cotton seed, 22, 26
cotton, 3, 22
cotton, cleaning, 150
cotton, color grade, 130, 281
cotton, cost, 28
cotton, cultivation, 23, 25, 26
cotton, dirty, 30
cotton, dye uptake, 166, 167
cotton, fiber attributes, 23–27, 30, 130
cotton, flattened tube, 23
cotton, harvesting, 24, 26
cotton, infestation, 25
cotton, irregular collapse of walls, 23
cotton, primary wall, 23
cotton, reflectance (Rd), 355, 404
cotton, secondary wall, 23
cotton, sticky, 25, 143, 175
cotton, warehouse categories, 391, 392
count & twist determined by market, 327
count spectrum, 304
count systems, 318
cover, 12, 15, 260, 371
creel blending, 129, 164
creel, 156, 168, 397, 398
creel, doubling, 164
creel, lap-winder, 159, 161
creel, power, 156
creel, roving bobbins, 175
creel, sliver combination, 164
creel, slivers, 216
creeping build motion, 177
crimp contraction, 369, 370
crimp recovery, 369, 371
crimp, fil, 12
crimp, ls, 206
crimped fiber, 78
cross link, 20, 21
crutching, 33
crystallinity, 42
crystallization rate, 46
cumulative draft, 72
cumulative frequency curves, 280
cuprammonium solvent, cellulose, 4
cushion of stock, 258
cushion roll, see cots
customer complaints, 129, 257, 277, 392
customer goodwill, 258, 306
cut tow, 48
cuticle, 31
cutting edges, 51
cutting, mill based, 13
CV of particular attributes, 130, 280, 314,
358, 390, 393, 411
cyclic errors, 73, 408
cyclone filters, 150, 348
CYMK system, 406
Cyros™, 362
dangers of extrapolation, 1
data obsolescence, 304
daughter tufts, 413
deburring, 219
defect levels, 17
defective fibers, 284
defining a blend, 390
definition of shear, 84
degradation of yarn quality, 81
delayed quality factors, 306
delivery speed, 276, 324
denier, 318
depreciation costs, 303
detergent scouring, 209
development of bulk, 384
development of machinery, 5
development of POY, 87
dew point, 344
dicotyledenous plants, 35
difference frequency, 398
differences between laydowns, 398
differences in fiber color, 284
differences within laydown, 398
differences, cost & price, 276
differential gearing, 169
differential shrinkage of fibers, 50
differential stress, 85
difficulties in processing, 47
digital data stream, 359
direct labor cost, 311
direct labor force, 309
direct number, 317, 318
direct winding, 242
dirt removal. tuft surface, 127
displacement of centers of production, 3
dissimilar bales in proximity, 400
dissolved gases, 344
distribution of pressure in nip zone, 81
distribution of the polymer flow, 44
diversity in m/cs, 205
divisions in testing, 350
doffer rail, 185
doffer, automatic, ring-frame, 185
doffer, grasping device, 185
469
470
Index
doffer, peg belt, 185
doffing, 142, 185
double creeling, 176
double fleeces, 34
doubled & mixing, 398
doubling in blending machines, 83
doubling & toothed drafting, 425
doubling mass constant, 424
doubling, 83
doubling, 83, 216, 398, 399, 422, 424
doubling, periodic errors, 399
down-twisting, 65
draft distribution, 414
draft magnitude, 411
draft ratio changes, 158
draft ratio of mill, 72
draft system, OE, 186
draft zone, 70
drafted fiber ribbon width, 418
drafting & doubling, 407
drafting errors, automatic control, 419
drafting or drawing, 71
drafting roll separation, 418
drafting system time constant, 419
drafting system, 70, 80, 84, 156, 168, 294
drafting systems, ls, size, 206
drafting theory, 407, 409–411
drafting waves, 78, 85, 282, 351, 410, 411
drafting zones, successive, 422
drafting, 14, 71, 74, 174, 321, 331, 408, 413–
415
drafting, added frictional forces, 79
drafting, aprons, 77, 79, 411, 415
drafting, CV at start-up, 415
drafting, damaged roll necks, 415
drafting, early staple processing, 83
drafting, error source identification, 413, 416
drafting, roll setting, 414
drafting, rotor spg, 413
drafting, setting, 84, 415
drafting, short-term errors, 412
drafting, sliver autonomy, 419
drafting, state of maintenance, 84
drafting, toothed, 83
drafting, uneven rubber hardness, 415
drafting, variance, 425
drafting, variance added, 422
drafting, variance, autonomous slivers, 419
drag coefficient, 439
draw zone, 156
drawframe & dedicated card in series, 398
drawframe autolevelers, 135
drawframe creel, 83
drawframe, 155, 156, 324, 325
drawframe, cross-feed, 398
drawframe, ls, 220
drawing & condensation, 83
drawing (high speed), start up, 47
drawing elements, 70, 75, 79, 49
drawing error, 164
drawing fil, 47, 75
drawing head, 155
drawing operation, 156
drawing POY, 42
drawing sliver, 75, 397, 77
drawing tow, 13
drawing, ls, front roller system, 217
drawing, m.m.fibers, 77
drawing, number of slivers in creel, 457
drawn fibers, 11
draw-off aprons, ls, 218
draw-off cylinders, ls, 218
draw-texturing, 42, 46, 102, 103
draw-texturing, crimp-resilience, 102
draw-texturing, degree of setting, 102
draw-texturing, economics, 102
draw-texturing, fil flats, 102
draw-texturing, logistical problems, 102
draw-texturing, sequential, 102
draw-texturing, simultaneous, 102
draw-texturing, speeds, 103
draw-texturing, surges, 103
draw-texturing, yarn properties, 102
dry air, slubs and fishes, 288
dry spg, 38
dry steam, 342
dryness fraction, 342, 343
ducting, fiber strings, 148
ducting, 148, 346
ducting, egress of dust and fiber
dust emission, 154, 156
dust filteration, 154
dust house, 150
dust removal, 117, 123
dust, fiber debris in air discharge, 150
dye affinity, 11, 21, 23, 78, 114, 284, 384
dye mandrels, 258
dye package, 234
dye sprigs (porous package centers), 234
dye streaks, 129
dye take-up see dye affinity
dyed fil yarns attributes 362
dyed knitted sleeves, 361, 372
dyeing and setting operations, 344
dyeing operation, 114
dyeing, autoclave, package density, 257
dyeing, errors, 115
dyeing, hank, 115
dyeing, knitted test sleeve, 114
dyeing, peg frame, 234
earthing equipment, 349
eccentric front roll, 408
economic effects of automating, 313
economic mill operation, 9, 301
economics, storage, low value bales, 391
economy changes, 277
edge crimp in staple yarn processing, 85
Index
edge crimping, 111
edge crimping, asymmetric quenching, 111
edge crimping, disoriented polymer, 111
edge crimping, quality control, 111
effect yarn, 19, 109
effective coefficient of friction, 59
effective electrode length, 356
effective fiber length, 78, 351, 352, 407
effects of shear, 84
efficacy of blending, 390
efficient light sources, 349
elastic & viscous forces, 383
elastic yarns, recovery properties, 369
elastomeric core, 112
electrical charges, 349
electrical energy, 342
electrified fibers coagulate, 349
electronic yarn boards, 362
electrostatic precipitators, 348
elongated plant cells, 23
elongation of specimen controlled, 364
elongational capabilities, 22
emerging fil cooled, 12
emerging fil drawn, 12
encoder, 418
end break, patrolling, 182
end break, roving frame, snow storm, 288
end breaks vs thins spots, 307
end breaks vs weak spots, 307
end breaks/cheese or cone, 315
end-break patterns, 291
end-break rate X length of down time, 307
end-break repair, 182
end-break, 69, 87, 174, 177–183, 183, 299,
308, 314, 314, 437, 442, 443, 448, 457
end-breaks & economics, 314
end-breaks & operator assignment, 307
end-breaks & quality, 298
end-breaks vs intolerable yarn faults, 315
ends breaks during patrol, 306
ends-down rate & defects/bobbin, 298
end-use, 20
energy balance in enclosed space, 346, 347
energy dissipation, 346, 347, 384
energy in joules, 342
energy input, electric motors, 346
energy removal from workspace, 347
energy source, bobbin, 177
energy, 341
English cotton system, 318
equalization of wages, 7
equation of fiber flow, 399
equipment maintenance, 424, 425
error amplitude, 407
error analysis, 294
error correlation, encoder, 292
error elongation, 408
error length limits, 362
error signal, 158
471
error
error
error
error
source, 293, 294
spectrum, 291, 408
wavelength from bale plucker, 395
wavelength, 73, 77, 114, 166, 399, 407,
408
error wavelengths carried fwd, 293, 294
error, bobbin to bobbin, 294
error, fiber-related, 294
errors in drafting, 414
errors, non-periodic, 414
errors, short wavelength, 83, 355
errors, start up, 419
errors, very long, 361
eveness tester, 283
evenness, 79, 358
evenness, sliver 155, 158, 299
exceptions reporting, 158
experimental plan, 281
expert system, 419
extensibility, 12
external forces on softened polymer, 383
extruder auger (screw}, 41
extruder barrel heating, 44
extruder die, debris at exit, 46
extruder screw, 42, 43
extruder spinneret, 44
extruder start-up, 44
extruder, 11, 12, 43
extruder, tapered metal screw, 44
extruders, ganged, 13
extruding polymer, 38
extrusion as drawing,71
extrusion thro’ same nozzle, 112
extrusion, pressure in barrel, 42
eye protection, 180
fabric appearance, 19, 22, 276, 282, 284, 295
fabric attributes, 18
fabric barré, 19, 78, 84, 96, 97, 106, 114, 129,
172, 175, 284, 294, 299, 355, 392, 393,
395, 398, 404, 417
fabric blotchiness or streakiness, 78
fabric cloudiness, 19, 414
fabric cover, 322, 369
fabric defects, demerit points, 289
fabric defects, see fabric faults
fabric durability, 276, 277
fabric faults, 289, 296
fabric filters, 348
fabric hand and appearance, 18, 101
fabric manf. problems, 254
fabric moiré, 294, 299, 414, 417
fabric patterning, 274, 275, 297
fabric shading, 114
fabric streaks, 114, 175, 284, 414
fabric weights calc, 321
fabric, burling & mending, 220
fabric, heat-set, 112
fabric, nature, 18
472
Index
fabric, seam pucker, 253
fabric, streaky, 274
fabric, woolen, felted, 220
fabric, woolen, milled, 220
false twist at rotor navel, 314
false twist spindle, 96
false twist spindle, suspension, 96
false twist spindle, tires, 96
false twist spindle, yarn drag, 96
false twist texturing, process stages, 384
false twist, 56, 62, 63, 92, 261
false-twist yarn, 384
false-twister, pin type, 94
fancy yarns, 19
fasciated yarn, false twist, 263
fasciated yarn, torque removal, 263
fasciated (wrapped) yarn, 263
fast Fourier transform (FFT), 166, 403
fault frequency, 283, 290
feed arrangements, 123
feed ribbons width, 264
feed roll damage, 425
feed systems, OE, 188
felting, 2
FFT analysis, 404
FFT of specific color wavelengths, 405
fiber-purchasing agent, 79
fiber & m/c interactions, 301
fiber, ls, damage, 206
fiber acceleration, 407
fiber accumulations, 123, 288
fiber acquisition policy, 281
fiber ageing, 47
fiber alignment, OE, 186
fiber array, 352
fiber attribute categories, 391
fiber attribute variability within flow, 164,
390, 392, 402
fiber attributes, 17, 50, 129, 280, 390, 394,
405
fiber avalanches, 420, 421
fiber batt, 117, 127
fiber blending, 48
fiber bonding along cuts, 51
fiber breakage during specimen prep, 353
fiber breakage, 28, 283, 413
fiber breaking stress, 353
fiber buckling, 104
fiber bulk, 50, 10
fiber buying policy, 301, 391, 392
fiber chokes, see chokes
fiber classification, 21
fiber cleaner, inclined, 127
fiber cleaning, OE, 186
fiber cleaning, re-entry into stream, 126
fiber clump division, 116, 117, 122, 390, 413,
421, 425
fiber clump flow, irregular, 425
fiber clump size, 122, 123, 124, 400
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
fiber
clumps, 117, 148, 412, 413
cohesion, 116, 413
color, 24
condensation as doubling, 117
condensation, 83, 117, 123
condensation, OE, 186
condenser, 75, 124
condenser, thin spots, 125
control, 262
control, chute feed, 134
control, error signal, 135
control, instability, 135
control, lag time, 135
control, linear density, delivery, 135
control, mechanical restraint, 79
control, on/off, 135
control, reserve box, 134
control, set point, 135
cost, 16, 301, 302
costs/yarn costs, 8
crimp, 4, 10, 116, 53
crimp, trailing ends, wool fiber, 382
damage, 128, 208, 221, 384
debond, 51
defects & drafting, 288
diameter variable, 205
distribution length, 397
dressings, 51
extent, 422, 423
feed systems, 117
fineness CVs, 166
fineness, 18, 19, 23, 24, 25, 280, 353
fineness, immaturity, 354
finish & treatments, 4, 47, 48, 51, 95,
116, 190, 191, 381
finish deposits, 288, 289
finish deterioration, 383, 384
finish, lighter fractions, 349
finish, ls, 206
finish, volatile fractions, 95
finishes as a size, 47
finishes formulations, 47
flats, 50, 383
fleeces, 117
flow control, 117, 133, 136, 407
flow, air turbulence, 148
flow, blockage, 133
flow, clumps, 399
flow, fractionation, 148
flow, shutdown cost, 133
flux, 455
handling, 128
helix angle, 323
homogenization, 117
hook removal, 81
hooks, 158, 189, 260, 353
in transit, specific volume, 123
laps, 174, 175
leading end, 407
Index
fiber length determination, laboratory, 279
fiber length variability, 293
fiber length, 23, 25, 59, 79, 205, 206, 279,
351
fiber length, sampling, cotton bales, 279
fiber loss, 182, 327
fiber lubrication, ls, 207
fiber mass, specific volume, 126
fiber migration & blend, 399–401
fiber migration in twist triangle, 85
fiber migration vs evenness, 86
fiber migration, 58, 75, 84, 262, 264, 375,
382, 384, 400, 411, 457
fiber migration, longitudinal, 164
fiber migration, yarn bulk, 375
fiber mixing process, 396
fiber mixture, step change, 395
fiber movement control, 216, 407
fiber movement, draft zone, 407
fiber movement, groups, 421
fiber oiling, 216
fiber orientation, 155, 161
fiber passing through fan blades, 128
fiber percentages in mixer, 396
fiber populations, 164, 421, 422
fiber prices, 6
fiber processing, 71
fiber pulse, 396
fiber purchase, see fiber buying
fiber quality control, 278
fiber reflectance, 284
fiber removal, layer by layer, 131
fiber restraining reaction, 410
fiber ribbon, optically measurement, 419
fiber rolls, accumulations, mill floor, 289
fiber selection, 312
fiber shuffling, 51
fiber slippage control, 56
fiber slippage, driving surfaces, 81
fiber stream, 123
fiber stream, specific volume, 122
fiber strength, 22, 25, 353, 388
fiber strings, 128
fiber structures, mechanics, 10
fiber tangles, 12
fiber tension, 374, 375
fiber testing, mass production basis, 392
fiber transfers, 87
fiber transport, 116, 123, 126, 128
fiber transport, OE, 186
fiber values in yarn, 391
fiber variability, 17
fiber yellowing, 281, 384
fiber, condensation, 123
fiber, consistent inventory in storage, 392
fiber, normalized flow, 396
fiber, physical characteristics, 114
fiber, re-cycled, 121
fiber, swollen & unswollen, 354
473
fiber, technical figures of merit, 391
fiber-crimp removal, 81
fiber-crimping, 50
fiber-finish level, 175
fiber-packing density variation, 381
fibers condensation, 128
fibers & fil, 18
fibers blend components lubricated separately,
211
fibers clumps, 390
fibers from moving zone, intermixing, 394
fibers migrate in drafting, 71
fibers swelling, caustic soda, 354
fibers, buckled, 50
fibers, carbon, 113
fibers, carpet, 10
fibers, detach from beard, 187
fibers, hooks, 158
fibers, short staple, 116
fibers, volume occupied, 10
fibrillation, 113
fibrils, 23
fibrogram, 352
fil curl, coil, or loop, 85
fil acceleration between godets,72
fil aspirator, 47
fil breaks, 46, 77, 96
fil changes due to age, 40
fil cooling before tension release, 49
fil cross-sectional shapes, 39
fil deformation, 89
fil drawing, 38, 42, 47, 114
fil dyeability, 4
fil flatting, 82, 384
fil lubrication, 47
fil minimum energy shapes, 384
fil processing, insufficient heating, 384
fil production, 11, 88
fil separation, 385
fil shrinkage, solidification, 39
fil spg speeds, 39
fil stability of POY, 42
fil strength, 42
fil take-off & drawing, 46
fil temperature control, 384
fil texturing process stages, 384, 385
fil texturing, edge-crimp, 85
fil texturing, torque, snarls & helices, 385
fil winding, 41
fil yarn production, 88
fil yarn, linear density, 368
fil yarns, 368
fil yarns, dye affinity, 282
fil yarns, polymer morphology, 282
fil yarns, twisted, 61
fil yarns, twisting & folding, 111
fil, aramid polymers, 113
fil, controlled conditions, 11
fil, dye affinity, 46
474
Index
fil, extruded, 11
fil, fibrillated, 113
fil, fibrillation process, 113
fil, glass, 113
fil, high-modulus, 113
fil, high-tensile, 113
fil, industrial, 113
fil, linear density, 45
fil, luster, 61
fil, minimum energy shape, 91
fil, multi-lobed, 113
fil, partially oriented (POY), 102
fil, setting, 90
fil, ultra-fine, 113
fil, winding, 90
fil, yellowing, 90
fil, 12
filtered polymer derivative, 40
filtration of flowing polymer, 38
filtration of signal output, 358
filtration, 348
finish & fiber deposits, 48, 114
finisher drawing, 159
finite element method of calculation, 440
fire and explosion risks, 348
fire code, 348
fixed costs, 308, 309, 312, 316
flammable airborne material, 348, 349
flatted fibers, constrained rotation, 383
flat-topped wire, 213
flat-waste, 287
flax (linum usitatissimum), 35
flax fiber fineness, 354
flax fibers, air permeability tester, 354
flax fibers, fiber division, 354
flax history, 1, 2
flax stricks, 231
flax use, 2, 35
flax, baling, 35
flax, bark removal, 35
flax, cells, 35
flax, curly top, 35
flax, fiber strength, 35
flax, hackled fiber stricks, 231
flax, hackling band, 231
flax, hackling tooth size, 231
flax, hackling, root end first, 231
flax, harvesting, 35
flax, linen cloth, 2
flax, lumen, 35
flax, natural drying, 35
flax, pulling machines, 35
flax, retting, 35
flax, scotching, 35
flax, seeds, 35
flax, stalks de-seeding, 35
flax, stooks, 35
flax, straw, 35
flax, wigwams, 35
flax, wilt (pathogenic fungi), 35
fleece of sheep, 31
floating fiber, 407, 410
flow charts, 13
flow of fiber passing thro’ reservoir, 395
fluted metal roll, 75, 125
fly deposited on textile material, 288
fly discharge, spg frame, 289
fly, 50, 156, 173, 282, 288, 346, 348
flyer speed, variable, 169
flyer twisting, 169
flyer, 168, 169
flyer, false twist, 169
flyer, presser arm, 169
flyer, roving support, 169
flyer, rubber grommet, 169
flyer, winding tension, 169
fog, 344
folding, 112, 220
folding, cable twisting, 111
folding, forming twist, 111
folding, see doubling
folding, torque balance, 111
force equation, rotor spg, 440
force equation, stationary model, 440
force equation, yarn balloon, 440
foreign fibers, 31, 120, 278
fork frame, 208
formation of necks, 77
fractionation of fibers, 162, 164, 390
freedom from faults, 276
frequency domain, 357, 358
friction spg, 186
friction twister, 98
friction twister, torque generated, 99
friction twisting, cumulative torque, 100
friction twisting, damage, drive rollers, 100
friction twisting, disk changes, 101
friction twisting, disk penetration, 100
friction twisting, disks pumping yarn, 100
friction twisting, drive surfaces, 100
friction twisting, fiber finish, 101
friction twisting, fil breakage, 100
friction twisting, filamentation, 101
friction twisting, grooved ball, 102
friction twisting, humidity, 101
friction twisting, limits to speed, 100, 101
friction twisting, productivity, 101
friction twisting, run-off angles, 100
friction twisting, run-on angles, 100
friction twisting, stacked disks, 100
friction twisting, thread line angles, 100
friction twisting, urethane surfaces, 100
friction twisting, yarn damage, 100
friction twisting, yarn tension, 100
friction, ring & traveler, 432
frictional bonds, 72
frictional restraints, 373
front roll nip, 407
Index
front roll velocity, 407
front roll, partial wrap of fibers, 174
fundamental wavelength, 291
funding, proceeds of sales, 309
g/9 km (denier), 318
gage pressure, 344
garnet, 221
gas equation, 343, 344
gear crimp, 53
gearing calc, 334
general mill expenses, 304
gilling (drawing, ls,), 216
gilling frames, productivity, 217
gilling, faller bars, 216
gin practice for cotton, 391
gin stand, 27, 28, 29
gin, ‘stick’ machines, 27, 28
gin, 24
gin, automatic feed control, 28
gin, bale presses, 28
gin, bales, 31
gin, boll traps, 28
gin, conveyor-distributor, 28
gin, disk-like saws, 29
gin, dryers, 28
gin, feed roller, 30
gin, feeder, 28
gin, fiber cleaning, 28, 29, 30
gin, fiber damage, 27, 29, 30, 31
gin, financial return, 28
gin, lint cleaners, 28, 31
gin, moisture content, 29, 30
gin, mote bars, 30
gin, out-turn, 30
gin, roller surface, 27
gin, saw blades, 27
gin, trash content, 31
ginning as a blender, 27
ginning byproducts, 27
ginning process, 25, 26, 117
glass fibers, 4
glass transition temperature, 20, 49, 381
goat hair, 31
godets, 46, 75, 82
godets, inclined pins, separate wraps, 82
Gossypium, 23
grains & yards, 318
greasy wool, 205
grids, 123
grin through, 16, 112, 260
growth rings, 23
guide pins, 75
hackling see flax, hackling
hand, 10, 22, 206
handling cost, 309–312
handling vs fixed costs, 313
hank length, 318
475
hard ends, 69, 288
hardening of rubber coverings, 73
harmonic analysis, 358, 408
harmonic errors, 73
harvesting, 23, 26
head-to-tail package connections, 258
heat energy, 341, 342
heat flow, 342
heat flow from newly extruded fil, 42
heat from m/c directly exhausted, 347
heat loss, conduction, 346
heat loss, radiation, 346
heat removal by water washing, 347
heat setting, 384
heat transfer, 39, 42, 346
heating system, 346
heat-setting of fil, 85
heat-stretching, 49
Heilman (French) comb, 217
helical minimum-energy shape, 385
hemp (cannabis sativa), 36
hemp cultivation, 36
hemp history, 2
hemp, fiber strength, 36
high draft, 176
high volume instrument (HVI), 279, 351, 355,
368, 392
high-capital-cost machines, 309
high-friction materials, 100
high-speed equipment, 6
high-speed texturing, 42
high-volume fans, 128
history, 1, 5, 119
HOK & productivity, 7, 9, 10
hollow spindle spg, 66, 106, 270
home furnishings market, 5
homogenizing multiple streams, 83, 131
honeydew, 25
horsepower, 342
hot pin, 77
hot-fluid texturing, 106
human exploitation, 5
human intervention, 9
humidity, workspace, 344
hump magnets, 126
Hunter scale of +b, 406
HVI calibration, 355
hydrogen bonding, 341
hysteresis, 371, 384, 385
image analyzers, 354
immature fibers, 284
immigrants carry their skills, 5
imperfect gas or vapor, 343
indirect labor force, 309
induction of lateral forces, 56
industrial case study, 165
industrial materials, 22
industrial practice, m.m. staple fiber, 391
476
Index
industrial revolution, 1, 5
industrial yarns, 250
information revolution, 1
input electrical energy to heat energy, 346
insect secretions, 25
insects, destructive, 25
inspection tours, 299
insulation properties, 89, 12
integrated values, 356
integration and automation, 86, 87
interest rates, 316
interference between helices & snarls, 384
interference, tube/balloon, 178
interfiber entanglement, 76
interfiber friction, 59, 85, 383, 384
interlocking structure, 58
international competition, 315
interpretation of data, 277
intersecting breaker bars, 50
intersecting faller bars, 216
inventiveness, 5
inventory control, 47
inventory storage, 258
investments in equipment, 8
irrecoverable low-grade heat, 347
irregular fiber flows, 77, 407
irregular input slivers, 81
irregular polymer flow in drawing, 77
irregular yarn faults, 357
irregularity index, 411, 423
irregularity or unevenness, 282
irreversible fiber migration, 399
joy riders, 184
justification of extra capital cost, 313
just-in-time (JIT) shipping, 258
jute & polypropylene, 2
jute (Corchorus), 22, 36
jute history, 2
jute substitutes, 36
jute, (abutilon theophrasti), 36
jute, bark, 36
jute, cultivation, 36
jute, fiber luster, 36
jute, fiber strength, 36
jute, overlapping cells, 36
jute, retting, 36
jute, stripping, 36
keratin, 31
labor cost/unit weight, 7
labor costs by process, 303
labor costs, 87, 118, 301, 303, 312
labor costs, ring spg vs rotor spg, 305
labor costs, spg vs count, 304
labor needs, 309
laboratory (or off-line) testing, 350
laboratory testing, 158, 351
lanolin, 208
lap quality, 161, 163
lap ribbon, doubling, 161
lap size, 163
lap winder speed, 161
lap winding, ss, 159
lappet guide, see pigtail guide
laps, split, 161
lap-up, 262
large package inside yarn balloon, 65
latent crimp, 92
latent heat, 342
lateral fiber migration, 85, 157, 260, 373
lateral fiber migration, see fiber migration
lattice apron, 117, 123, 125
lattice card feed, 213, 223
lay gear, 65
laydown pattern, 392
laydown variations, 129
laydown-laydown variation, 389
lea strength = count-strength product, 368
leading & trailing hooks, 353
lean yarns, 262
length data series by choice, 362
length variable, 205
length/mass, 317
length-biased sample, 280
lighting, good reflectors, 349
limited total fiber denier, 49
linear density CV/theoretical values, 421
linear density errors, 400
linear density transducer, 157
linear density variance among strands, 422
linear density vs time elapsed, 356
linear density, 24, 57, 58, 317, 321, 324
linear density, capacitance of strand, 356
linear velocity of yarn, 57
linen thread, plying, 232
linkage, 155
linked spg, 118, 185, 255
linking & automation, 87
lint & linters, 24
lint and harmful dust removal, 348
lint collection systems, 174
liquefying by melting, 38
liquefying by solvents, 38
liquefying cellulose, 4
liquid flow rate, 45
liquor degradation, 208
liquor evaporation, 212
liquor flocculation, 212
liquor make-up, 208
liquor pH, 208
liquor temperature, 208
liquor troughs (or bowls), 208
load cell, 353, 364
load disribution among fibers, 375
load distribution between feeds, 82
long fibers, 15, 205, 207
Index
long knitted dyed sleeve, color testing, 372
long staple, cutting tow, 51
long-chain molecules, 4, 20
longitudinal fiber migration, 85, 86, 165, 399,
400
long-staple combing, 217
long-staple roving & spg, 218
long-staple spg, 205, 218
long-staple yarn production, 15
long-term errors, 83, 118
long-term variables, 310
loss of production, 182, 307
lost production vs pneumafil levels, 308
lower portion of balloon, 444
lumen, 22
m.m. carpet yarns, 15
m.m. fiber bales, 216
m.m. fibers, 22, 38
m.m. fibers, history, 4
m.m. fibers, oligomers, 157
m.m. fibers, process r.h., 212
m.m. fil, 4
m.m. staple fiber production, 48
m.m. staple fiber, 4
m.m. fiber damage, 283
m.m. fiber, sampling procedures, 279
m.m. fibers, 8
m.m. fibers, drawn fiber, 284
m.m. fibers, finish concentrations, 284
m.m. fibers, oligomers, 284
m/c clogging, 128
m/c component speed calc, 339
m/c created errors, 407
m/c design, 309
m/c dimensions/fiber length, 218
m/c element eccentricities, 282
m/c element velocity, 321
m/c element, abrasion, 123
m/c element, maintenance, 123
m/c idle, 313
m/c maintenance, 174, 299
m/c productivity calcs, 324
m/c setting, 128, 299
m/c speeds calcs, 334, 335
m/c tooth size, 122
m/c wear, 283
m/c, OE, sliver input, 186
magnet to remove ferrous material, 126
maintenance costs, 304
maintenance cycles, 155
maintenance of cutter, 51
maintenance of roll settings, 79
maintenance staff costs, 308
management costs, 304
management of repairs, 291
market size, 15
markets, spun yarns, 9
married fibers, 46, 282
mass constant, 397, 398
mass distribution unevenness, 399
mass flow control by passive devices, 70
mass flow, 321
mass testing of cotton bales, 279
mass transfer, 39
mass, 341
mass/length, 317
matching productivities, 324
materials handling, 9
measurement of twist, 359
measurements on staple fibers, 351
measuring r.h., 346
mechanical & thermal history, 114
mechanical cleaning function, 117
mechanical draft ratio, 321
mechanical energy, 342
mechanical errors, 408
mechanical feeds, 125
mechanical properties, 10
mechanical working of melt, 43
mechanics of false twist, 62
mechanics of fiber wraps, 70
melting point (Tm), 85
melt-spg, 41
merge changes, 13, 392
meshing eccentric rolls, 408
metallic (saw-toothed) wire, 213
metering device, 45
metering pump leakages, 45
microbial pathogens, 32
micronaire index, 24, 280, 353, 354, 403
micronaire variations, dye affinity, 404
micronaire, error wavelength, 166
microscopic examination, 354
migration of solvent, 40
migration of the textile industry, 8
mill application of incoming bales, 392
mill as mixer, 389
mill balance, 183, 184, 313, 327
mill environment, 283, 346
mill management reports, accuracy, 307
mill performance, 310
mill pipeline, 397
mill processing, 17, 344, 345, 397
mill productivity target, 325
mill testing of incoming bales, 392
mill, fiber flow branches, 397
mineral & vegetable particles, 125
minimizing mill imbalance, 314
minimum cost, 316
minimum energy condition, 385
minimum irregularity, 423
minimum processing trouble, 350
mixed noil, 216
mixers, dispersion of fiber flow, 395
mixing machines, 395
mixing volume, 395, 397, 400
mixing within machines, 116
477
478
Index
modacrylics, 21
modified skein shrinkage test, 371
modified twist, 261
modified yarn structures, 260
mohair, 31
moiré effects, 114
moist air, 2, 344
moisture content of fibers, 212, 356
molecular dislocations, 85
molecular orientation, 42
molecular structure, polymers, 11, 12, 38
monitoring, 11, 158, 350
monitoring, computer linked, 158
Morel beaters, 213
Morel clearance, 213
morphological structure, 42, 114
moth & beetle larvae, 31
mothproofing, 32
moving interval, 166
moving zone variation, 400
moving zone, bale laydown, 394
moving zone, horizontal slices, 394
mule spg, 218, 230
multi-function fiber measurement, 368
multiple doublings, 205
multiple drawings, 424
multiple parallel fiber flow paths, 324
multiple weak spots, 413
multiple, consecutive draw zones, 72
narcotics & fiber, 2
natural draw ratio, 47, 77
natural fibers, 22, 393
natural polymers, 20
navels, ceramic, 195
neck position, stabilize, 77
needle melts, 16
nematode, 25
nep control, 147
nep creation, 25, 123, 128, 458
nep, 10, 19, 23, 281, 283, 284, 285
nep, OE combing roll, 285
neps in card web, 281
neutralization of acid, 41
new technology, 303
Nickerson-Hunter colorimeter, 130
nip line oscillation, 408
nitrocellulose fil, 4
noil, 151, 162, 163, 287, 390
noise level, 181, 216
non-fibrous material, 284
non-lint material, 122, 150
non-oxidizing atmosphere, 384
non-reworkable waste, disposal, 150
non-wovens, 22
normalized productivity. see HOK, 7
normalized variance, 381
noxious chemicals, 348, 349
nubs, 19
number of bobbins w piecings, 315
number of m/cs needed, 324, 325, 327, 328
number of piecings, 315
number of spinnerets, 13
numbering ply yarns, 319
nylon, 21
obliquity curve, 58
OE history, 186
offtake passage over a bale, 396
OHP, 304, 308
OHP/count, 306
oiled wool fibers, coeff friction, 381
oily soiling, 4
oligomers accumulations, 349
on line monitoring, capital cost, 291
on-line monitoring yarn defects, 351
on-line monitoring yarn hairiness, 351
on-line monitoring, 290, 291, 300, 350, 351,
358
on-line sampling, 362
open-end (OE) spg, 67, 185
open-end spg, brief history, 453, 454
opening & carding, trash removal, 283
opening & cleaning, 125, 153
opening line & carding, 425
opening line layout, 119
opening line, 116, 122, 153, 310
opening line, mixing, 394
opening of new markets, 5
opening, 117, 118
operation without undue disruption, 50
operational factors, 313
operational flexibility, 117
operational phases, 118
operator assignment & OHP, 309
operator assignment, 299, 306, 307
operator efficiency, 299
operator hrs/kg product (HOK), 303
operator hrs/lb product (OHP), 303
operator hrs/task, 303
operator patrolling, 306
operator perception of delayed quality factors,
306
operator productivity, 304, 305
operator productivity, history, 303
operator training, 298, 299
operator wage rate, 306
optical character, 355
optical masks, 363
optical testing of flowing strand, 356
optimum cost, 312, 313
orbital movement, 70
order book, 314
organochlorine compounds, 32
orientation, 70
oscillating rolls, 68
oscillation of the neck, 78
overall draft, 72, 321
Index
overall draft ratio, OE, 188
overdrawing, 424
overfeeding, 92
overhead costs, 301, 304
overhead rail systems, 184
oversupply, 4
package build, 181, 235
package chains in feed, 258
package density, 234, 235, 257
package dyeing, 234
package moisture content, 257
package shoulders, 239
package size, 250
package size, OE, 186
package storage, 234
package structure, 234, 235
package structure, periodic change, 237
package transport, 234
package, cross-wound, 234, 235, 237
package, density variation, 238
package, inter-yarn forces, 237
package, ribboning, 238, 239
package, stability, 237
package, traverse motions/revolution, 239
package, unwinding, 234, 257
package, yarn interlace, 237
papilla, 31
parallel portions of yarn loaded, 366
partial pressure of moisture in air, 344
partial pressures, 343
partially oriented yarn (POY), 42, 47, 102
partial-wrap design to improve grip, 81
particulate level in air, monitoring, 349
passage of drawing, 76, 159
pathogenic fungi, 25
patrol events, 253
patrolling sensor, 291
payback time, 309
pendulum tester, 368
perfect commodities, 8
perforated apron, 262
perforated hollow front roll, 262
performance correlations, 307
performance of machine in serial line, 293
performance vs spindle assignments, 306
periodic errors of random phase, 398
periodic shearing, 33
periodic variations, 76, 114, 115, 281, 397,
398, 408, 425
periodogram, 358, 402
permeability, 12
personnel costs, 308, 310
Peruvian Tanguis, 23
pesticide, 25
pests, 23
phase change of blend components, 400
phases, OE, 186
phasing errors, 398
479
physical basis of texturing, 89
piecer, mechanical, 185
piecing 1 for each bobbin used, 315
piecing, 179, 182, 282, 283, 287, 309, 310,
315
piecing, automatic, 190
piecing, fault source diagnosis, 287
pigtail guide, 177, 178, 268, 427, 446
pigtail guide, rotating, 446
pilling, 4, 388
Pima cotton, 23
pin drafters, 216
pin twisting, economic limitations, 98
pinned feed systems, 208
pinned rolls, 125
pinning density, 216
pin-twister machine, limitations, 96
pin-twister, 93, 97
planning, 9
plied ply yarns, 61
plied yarns, 253
ply direction, 60
ply twist, 60
ply yarn numbering, 332
ply yarns, 250
Plyfil system, 265
Plyfil yarns, index of irregularity, 270
Plyfil yarns, twist balance, 270
plying (doubling), 65
plying 2 fold yarns, 251
plying by two-for-one twisting, 65
plying, unwanted three-fold yarn, 251
pneumafil collection, 175, 182, 308, 346
pneumafil devices, 262
pneumafil production, 291, 416
pneumafil waste, 150, 287
pneumafil, fiber crimp & elongation, 416
pneumatic separators, 126
pneumatic trumpets, 417
polyacrylonitrile fibers, 38
polyamide fil, 4
polyamides, 21
polyester, 21
polyester/cotton, 5
polyester/wool, 5
polyethylene, 22
polymer cross-linking, 44
polymer debris, 282
polymer degradation, 42
polymer derivative ripening, 40
polymer discoloration, 38, 384
polymer drip, 45, 282
polymer feed, continuous molten, 42
polymer filtering & pumping, 41
polymer filtration, 43, 45
polymer melted in barrel, 12, 41
polymer metering, 12
polymer morphology & dyeing, 114
polymer morphology, 84, 277
480
Index
polymer over-heating, 42
polymer oxidation, 38
polymer pressurize, 12
polymer pumps, 42
polymer solidification, 38
polymer solution, 40
polymer supply, 12
polymer transport, 41
polymer viscosity, 41, 42, 45
polymer, liquefying, 38
polymer, orientation, 38
polymer, visco-elastic properties, 85
polymeric materials, 20
polymeric structure, elongation, 112
polymerization/spg systems, 42
polyolefins, 22
polypeptide, 31
polypropylene, 22
polypropylene, UV degradation, 113
polyurethanes, 22
poor cleaning procedures, 288
poor spg, poor winding, 315
portable hygrometer, 346
post-spg processes, 234
power cost, 301, 303, 311, 312, 342
POY, 47, 102
poymer filtering, 12
precision winding, 242
premature fiber acceleration, 81
preparation costs, 304
pressure of the steam, 344
pressure on inner fibers, 57
pressure on stationary cushion rolls, 73
prickle, 10
probability distribution of tenacity, 314
probability distribution, applied stress, 314
probability of piecings, 315
process costs, 212
process efficiency, 324
process integration, 86, 87
process of elongation, 71
processing error, 281
processing variations, 390
processing, ls, large diameter rolls, 206
product delivery, 8
product judgment in dyed state, 393
product judgment in the greige state, 393
product labeling, 300
product price, 306
product sampling, 300
production calcs, 336–338
production of polymers, 9
production of viscose rayon yarn, 4
production speeds, 42
production systems, 10
production, 24 hr/day, 155
production, m.m.fibers, 38
productivity gains, 1
productivity, 15
productivity, 303
productivity, normalized, 303
productivity, opening line, 324, 326
productivity/spindle, 308, 309, 327
profit (or loss), 301
profit margins, 276
prompt delivery, 258, 316
proper ventilation, 349
prosperity variable, 1
prudent recycling, 287
psychrometric charts, 344
publicly financed companies, 5
pumps, 38
purpose of twist, 56
purposes of drawing, 70
purposes of texturing, 89
quality & price, 276
quality & quality control, 276
quality and economics, 306
quality assurance, customer protection, 287
quality audits, 159
quality control & testing, 350
quality control program, 278
quality control, 17, 185, 276, 350
quality control, intermediate products, 281
quality deficiencies, market, 306
quality factors, 278
quality of product, 8, 315
quality, determined by the user, 276
quality, fabric attributes, 276
quality, fiber attributes, 276
quality, yarn attributes, 276
quasi-random variations, 71, 403
quasi-stationary balloon, 439
quench, unequal cooling, 46
quenching, 46
r.h., 344
radial compression, 373
radius of curvature, on yarn surface, 374
rag picker (shredding), 221
ramie degumming, 36
ramie (B. Tenacissema), 36
ramie (Boehmeria Niva), 36
ramie decortication, 36
ramie, fiber removal from stalks, 36
ramie, fiber strength, 36
ramie, thick-walled cells, 36
random error, 76, 83, 293, 389, 408, 425
random speed varying devices, 109
rapid technological change, 3
rare-event stoppages, 155
ratch setting, 49, 51, 206
rates of improvement, 303
ratio of linear densities, 321
real & false twist, 61
real blend variation, 402
reason for testing, 350
Index
re-break long fibers, 50
re-condensation, 67
recording mass variations, 358
rectilinear motions, 162
recycled blended fibers, 286
recycled polymers, 286
recycling waste, max percentage, 133
recycling, fiber identification, 221
recycling, limited by regulation, 221
refrigerated air-conditioning system, 347
regenerated fibers, 20
regional balance of technology, 304
regularity, see evenness
reinforcement of composites, 113
relative conditions undergo change, 316
relative costs, 302
relative movement of clamps, 353
relative movement of segments of molecules,
85
relative rotational speeds, 177
reliability, 316
remnants of sliver & roving, 150
reputation, 276
research, 8
residual fault rate, 287
residual twists, 384
residual vegetable matter, 210
residues of salt, calcined, 212
resultant forces acting on fibers, 57
return air systems, 346
returned shipments, 258
review of processes, 16
re-workable waste, 151, 392, 416
Reynolds number, 438, 439
ribbon breaking, package lift, 236
ribbon breaking, package oscillation, 236
ribbon breaking, slippage, 236
ribbon breaking, variable drive speed, 236
ribbon lap machine, 161
ring & traveler, 65, 178
ring & traveler, force analysis, 433
ring & traveler, micro-welds, 432
ring & traveler, run-in, 433
ring & traveler, sliding track, 432
ring & traveler, wear, 433
ring bobbin wind, 243
ring bobbins, temporary storage, 450
ring bobbins, unwinding, 450
ring burn, 442
ring damage, 180, 442
ring flange, 178 , 179
ring frame productivity, 327
ring frame, monitoring, 291
ring frame, winder head balance, 255
ring life, 443
ring rail lift, 178
ring rail, 177
ring run in, 180
ring size, 182, 183
ring spg, twist triangle, see twist triangle
ring spg, 168
ring spg, costs, 304
ring spg, forces acting, 181
ring spg, m/c, 205, 218, 230
ring spg, real twist, 62
ring spg, yarn above pigtail guide, 433
ring spindle productivity, 177
ring tube storage, 180
ring tube, 175, 234
ring tube, package structure, 451
ring yarn packing density, 381
ring yarn, self-locking structure, 85
ring, lubrication, 179
ring, micro-welds, 179
ring, run-in, 179
ringframe limitations, 182
ring-rail, 65
rings centered, 179
ring-spg m/c, 175
ring-spun yarn, structure, 260
rise in wage costs, 6
risk & cost, 258
roll ageing, 175
roll buffing, 156, 174, 416
roll circumference/fiber length, 174
roll damage, 49
roll eccentricity, 291, 408, 415
roll laps, 206, 349
roll layouts, 80
roll pairs, 125
roll separation transducer, 420
roll setting, 78, 420
roll size, 175
roll squeeze on soft fil, 82
roll weighting, 174, 175, 217
roller drafting system, 73, 171, 407, 414
roller drafting, doubling, 422
roller errors, 73, 76
roller gin, 30
roller or godet defects, 76
rolls, deposits of oil or grease, 175
ropes & cables, 113
rotary gills, 216
rotating condenser, 117, 123
rotating rings, 179
rotating sliver passage, 157
rotating spikes or teeth, 121
rotational speed of twister, 57
rotor bearing systems, 196
rotor cleaning, 190, 198
rotor collects fibers, 67
rotor deposits, 192
rotor doubling, 425
rotor fiber flow, 192
rotor groove lapped by rotating yarn, 193
rotor groove, 193
rotor groove, fiber ring, 193, 198
rotor input, fiber flow, 186
481
482
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
Index
m/c, productivity, 306
piecing, diagrams, 199
piecing, end conditioning, 198
piecing, introduce starter yarn, 198
piecing, sliver end conditioning, 199
piecing, start up, 198
piecing, timing, 198
productivity, 328
productivity, OE, 67
spg, 9, 56, 186, 347, 425
spg, abrasive fiber or dust, 203
spg, acrylic fibers, 202
spg, air bearings, 196
spg, airflow, 189, 192
spg, assembly of fibers in rotor, 192
spg, automatic doffing, 198
spg, automatic piecers, 187, 305
spg, automatic rotor cleaning, 198
spg, automatic start-up, 198
spg, blend evenness, 456
spg, blend yarns, 202
spg, bridging fibers, 200, 459
spg, built-in monitoring, 198
spg, bunch winder, 198
spg, capital cost /lb yarn, 187
spg, capital cost, 186
spg, centrifugal force, 193
spg, channel damage, 193
spg, chokes, 191
spg, cleaning aperture, 192
spg, cleaning capability, 189
spg, cleaning cycle, 191
spg, cleaning edge, 188, 189, 192
spg, combing roll housing, 192
spg, combing-roll clothing, 457
spg, conical transfer surface, 459
spg, cotton fibers, 190
spg, cranked doffer tube, 462, 463
spg, CV of linear density, 456
spg, debris in rotor, 190
spg, dirt box, 189
spg, doffing tube, 193
spg, doublings, 455 , 456
spg, end-breakage, 196, 201
spg, equipment maintenance, 456
spg, false twist, 459, 462
spg, feed roll & plate, 457
spg, fiber assembly, 459
spg, fiber breakage, 457
spg, fiber cleaning, 191, 202
spg, fiber crimp, 201
spg, fiber damage, 189
spg, fiber extent, 193
spg, fiber fineness, 201
spg, fiber finish, 201, 202
spg, fiber flux, cyclic variations, 457
spg, fiber length, 200, 201
spg, fiber peeling point, 459
spg, fiber requirements, 200
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
rotor
spg, fiber tenacity, 201, 202
spg, fiber-transport channel, 192
spg, financial return, 191
spg, glazed belt, 191
spg, hooked fibers, 193
spg, hot rotor, 192
spg, incipient yarn in rotor V, 461
spg, ingoing fiber/outgoing yarn
interfere, 193
spg, large yarn packages, 197
spg, load carrying ability of wrappers,
459
spg, long-term errors, 457
spg, low-friction finishes, 201
spg, m.m.fiber, 189, 190, 201
spg, m/c element abrasion, 189
spg, m/c element operating lives, 203
spg, m/c maintenance, 203
spg, m/c transmissions, 197
spg, navel, 194–196
spg, nep production, 190
spg, noil, yarn quality, 201
spg, number of fiber in X section, 456
spg, oligomer, 202
spg, organized errors, 457
spg, overheated whorl, 191
spg, piecer control, start up, 200
spg, piecing, 186, 198
spg, Poisson distribution, 456
spg, preparation errors, 457
spg, prepared starter tubes, 198
spg, random variations, 457
spg, real + false twist, 194
spg, recognition of end breaks, 462
spg, retention of twist in rotor, 462
spg, rotating navel, 194
spg, rotor debris, 191
spg, rounded groove, 462
spg, single tape drive, 191
spg, sliding contact, 193
spg, sliding path of a fiber, 459
spg, sliver feed, 189
spg, sliver loop in feed, 191
spg, sliver weight, 189
spg, slivers preparation, 456
spg, speed, 186
spg, spin limit, 201
spg, standard deviation, 456
spg, starter-yarn bunch, 198
spg, start-up devices, 200
spg, stationary navel, 194
spg, suitable fibers, 188
spg, tapered duct exit, 193
spg, transfer channel, see transfer duct
spg, transfer duct, 189, 193
spg, transfer yarn tails, 198
spg, transient twist depletion, 462
spg, transport of full bobbins, 198
spg, twist multiple, 195, 201
Index
rotor spg, twist propagation, 461
rotor spg, twist surges, 462
rotor spg, twist trap, 194
rotor spg, twist traps, 462
rotor spg, vector addition, 457
rotor spg, waste fibers, 202
rotor spg, wedge action of V groove, 193
rotor spg, winding, 197
rotor spg, wrapper fiber production, 459
rotor spg, yarn count, 202
rotor spg, yarn formation, 192
rotor spg, yarn handling costs, 197
rotor spg, yarn package, 197
rotor spg, yarn rolls in navel bore, 194
rotor spg, yarn take-up, 197
rotor spg, yarn torque, peeling point, 461
rotor spg, yarn withdrawal tube, 462
rotor suction, 189
rotor supports fibers, 67
rotor yarn packing density, 381
rotor yarn structure, 361, 378, 379, 381, 460
rotor yarn, 378, 459
rotor yarn, bridging fiber, 460
rotor yarn, bulkiness, 379
rotor yarn, calculated machine twist, 379
rotor yarn, ceramic guide, 381
rotor yarn, end-breakage, 189
rotor yarn, evenness, 189
rotor yarn, false twist, 460
rotor yarn, fiber helix angles, 378
rotor yarn, geometry of navel, 461
rotor yarn, hairiness, 189, 379
rotor yarn, nep production, 189
rotor yarn, projected fiber length, 460
rotor yarn, rotating navel, 460
rotor yarn, rotor V, 461
rotor yarn, tenacity, 189
rotor yarn, twist distribution, 460
rotor yarn, twist levels, 378
rotor yarn, untwisting, 379
rotor yarn, yarn waists, 460
rotor yarns, calculated twist, 360
rotor, air r.h. inside rotor, 192
rotor, air stream, 187
rotor, draft ratio at feed, 187
rotor, fiber condensation, 187
rotor, fiber sliding path, 193
rotor, OE, 67
rotor, real yarn twist creation, 193
rotor, sliding wall, 193
rotor, yarn twist runs back, 193
rounding results, 325, 327
roundness of drafting elements, 73
roving bobbin rotation, air pump, 289
roving costs, 304
roving count, 172
roving draft, 173
roving frame, 168
roving m/c, 218
483
roving m/c, bobbin-lead, 169, 171
roving m/c, break out supply, 173
roving m/c, end breaks, 173
roving m/c, fiber discharge, 173
roving m/c, fly, 173
roving m/c, flyer-lead, 169, 171
roving m/c, grommet, 172
roving m/c, headstock, 171
roving m/c, lay gear, 171
roving m/c, relative winding speed, 171
roving m/c, shut down, 173
roving m/c, size, 171
roving m/c, snowstorm, 173
roving m/c, spindle rows, 172
roving m/c, traverse, 171
roving m/c, traversing blowers, 173
roving m/c, traversing suction nozzles, 173
roving m/c, winding-on speed, 171
roving package damage, 169
roving productivity, 56, 325
roving quality, 299
roving size, 171
roving stop mechanism, 291
roving stop system, 182
roving testing, 281
roving twist, 69, 168, 169, 171, 173
roving, bobbin density, 169
roving, end breaks, 171
roving, hard ends, 173
roving, package, 171
roving, twist set, 173
roving, variable wind-on speed, 169
roving, winding tension, 169
rubber cots, flats, 415, 419
rubber covered roll, 75, 157, 216
rubber hardness, 174
running speeds, coeff of friction, 381
rust, 344
safety, 119, 129, 153, 154
sales & price, 8
sales system inertia, 258
sales transaction chains, 8
sample conditioning, 350
samples, 84, 276, 279, 350, 390, 422
sampling frequency, 287
sandwich blender, 131, 132
saw teeth, 127
scanning laser, yarn arrays, 372
schappe, (spun silk), 38
scouring liquor, 208
scouring, 15
Sea Island cotton, 23
seasonal changes, 392
seed & non-lint material, 30
seed coat fragments, 125
segmented polyurethane, 112
segments of long-chain molecules, 84
segregation, non-lint & usable fiber, 150
484
Index
self twist (ST) yarn structure, 379, 380
self twist m/c, 271, 273, 274
self-twist m/c, costs, 274
self-twist m/c, oscillating front roll, 274
self-twist m/c, productivity, 274
self-twist m/c, roller-drafting, 274
self-twist m/c, shuffling/twisting rolls, 274
self-twist spg, 271
self-twist spg, plied worsted rovings, 271
semi-worsted systems, 219
sensor transport systems, 291
sensors in machine, 350
sensors, 350
sensors, insensitivity to environment, 350
separation of fibers, 425
separation of winding and twisting, 65
separator plates, 177
sequential flow of batches, 424
service, 8, 276
settling tanks, 209
severe drafting action, 67
sewing m/c, needle melts, 252
sewing thread, cotton, gassed, 251
sewing thread, cotton, singed, 251, 252
sewing thread, 16, 250, 251, 252
sewing thread, cotton, plied, 252
sewing thread, dye affinity, 253
sewing thread, guide clogging, 252
sewing thread, linen, 252
sewing thread, mercerized, 252
sewing thread, needle eye clogs, 252
sewing thread, Nomex, 252
sewing thread, silk, 252
sewing thread, waxed, 252
sewing thread, yarn hairiness, 252
shear effects, blend, 85, 400
shear rate in extrusion zone, 42
shearing of two pinned surfaces, 207
sheep breeds, 33
sheep confined for clip, 33
shelf-life, 48
shipping costs, 8, 258, 316
short fiber acceleration, 410
short fiber content, 164, 353, 417
short fiber removal, 59, 161
short mechanical process, 11, 87, 303
short staple, cutting tow, 51
short-fiber content, 280, 281, 308, 403
shorthand twist notation, 60
short-staple cutter, 52
short-staple yarn production, 14
shrinkage, 321, 381, 384
shuffling rolls, 68
signals for information, 350
signals, outliers, 359
silk & denier, 37
silk & man-made fibers, 2
silk ‘books’, 37
silk attributes, 2, 37
silk cultivation, 37
silk glands of larvae, 37
silk harvesting, 37
silk history, 1
silk reeling, 37, 38, 113
silk, Bombyx, 37
silk, cottage industry, 113
silk, fibroin (amphoteric colloid), 37
silk, fibroin extrusion, 37
silk, fibroin ripening, 37
silk, fil, 37
silk, filature, 113
silk, gum as a size, 113
silk, hanks, 113
silk, long-chain molecules, 37
silk, sericin (natural gum), 37
silk, spg, 38, 113
silk, staple fiber, 37, 113
silk, throwing see silk spg
silk, yarn production, 37
silk, yellowing, 37
silkworm (lepidoptera), 37
simple draft calculation, 320, 321
singeing, 251, 252, 349
single-end tests, 365
sinusoidal blend error, 400
Sirospun spg, 219, 268, 269
Sirospun spg, cost, 269
Sirospun yarn, 270
Sirospun yarn, twist balance, 270
Sirospun yarn, weavability, 268
sisal & manila hemps, 36
skein dyeing, 236
skein shrinkage, 371
skein testing, 366
skeins weights, 355
skirting, 33
Sl system, 341
slack fibers buckle, 375
slippage between rolls & fiber, 81
slippage, 321
sliver attributes, 76
sliver ball, 217
sliver can, 148, 149, 155, 156, 159
sliver can, crushed sliver, 149
sliver can, piston, 149
sliver cohesion, 75
sliver condensation, 157
sliver condenser, 156
sliver CV, fiber attributes, 84, 161, 404
sliver doubling, 156
sliver drafting, 156
sliver handling, 184
sliver lapper, 161
sliver output, 118
sliver preparation, 155
sliver ribbon, 161
sliver sampling, 158
sliver storage, 191, 148
Index
sliver tension, 161
sliver testing, 281
sliver trumpet, 157, 418
sliver, 14, 117, 148
sliver, carded, 155
sliver, color striations, 406
sliver, combed, 155
sliver, ls, samples, transport, 281
sliver, mass sensors, 217
sliver, over-worked, 158
slivers, roll grip variation, 161
slow adoption of new technology, 313
slub creation, 283, 420
slubbing (roving), 216
slubbing, cheeses on mandrel, 230
slubbing, drafting, 230
slubbing, winding, 230
slubbings, rubbed for fiber cohesion, 229
slubs, 69, 288, 361, 421
slubs, periodic outbreaks, 421
sludge centrifuges, 209
sludge disposal, 212
smoothing length, 389
smoothing, drawframe creel, 389
snagging & pilling, 388
socialism, 6
softening point (Tg), 85
soiled wool, 33
soluble organic salts, 212
solvent distributions, 40
solvent removal, 39
solvent scouring, 209
sources of defects, 282, 288, 349, 399
space costs, 301
span length, 352, 353
spark-free motors, 348
specific enthalpy, 343
specific heat, 341, 342
specific volume of steam, 344
spectrogram array, 292, 294
spectrogram, 73, 158, 291–293, 296, 356–358,
408, 420
spectrum of sinusoidal variations, 399
spectrum of wage rates, 7
speed frame (roving frame), 216
spg bobbins, mixing, 294
spg cost vs yarn count, 305
spg efficiency, 300
spg end-breaks, 289, 298, 299
spg faults produced/hr, 300
spg faults, 278
spg frames, 15
spg m/c substitution, 312
spg productivity, 314
spg tension, 178
spg, cotton, 12
spg, individualized fibers, OE, 67
spg, long-staple, 12
spg, process, OE, 185
485
spg, short-staple, 12
spg, ss, 168
spg, wool, 12
spiky trash, 127
spin finish, 41, 47
spin limit, 316
spindle assignment, 253, 298, 308
spindle eccentricity, 180, 181
spindle life, 181
spindle productivity, 183
spindle speed, 176, 180, 181, 183
spindles required, 183
spindles/machine, 327
spinline, 11
spinneret blockage, 45
spinneret die, 45
spinneret, 12, 38, 39
spinners assignment, see spindle assignment
spinner’s costs, 234
spinpacks, 42
spiral cutter, 51
splice tails, 288
splicing chamber, 248
squeeze rollers, 208
stable twist in flowing yarn, 63
stages of drafting, 321
stages of drawing, 216
stages of processing, 6
stages of production, chain reactions, 407
standard fabric thickness tester, 371
standard of quality, 277
standard tests, 279
standards of living, 7
staple fibers, 18, 294
staple fibers, man-made, 11
staple length, 59
staple vs fil, 22
staple yarn production, 13
staple yarn structures, 373
staple, drawing, 14
staple-yarn manufacture complex, 11
staple-yarn systems, 11
static electricity, 4, 47, 174
static loading, 75
statistical control techniques, 355
steam pressure & temperatures, 341
steam properties, 342
stepless drive motors, 217
stiffness & bulk, 10
stiffness of fiber loops, 18
stiffness of hairs, 18
stock control, 79
storage, r.h., 216
strand flow thro’ a torque zone, 62
strand(s) wrapped around a core yarn, 66
strap & bale covers removal, 120
streakiness, 21, 129
stream orientation, 46
strength of POY, 42
486
Index
strength of a twisted bundle, 58
stress decay, 371
stress removal by polymer softening, 385
stress-free, helical condition, 385
stretch breaking for m.m. fiber, 13, 217
stretch yarn, 369, 371, 385
stretch, 12
stretch-break, mill based, 13
stretch-breaking system, 15, 48 49, 50
stringing-up, 95
strip charts, 356, 357
stroboscope, 175
stuffer box, 50, 53, 105
stuffer-box texturing, 104, 105, 106
stuffer-box yarns, plied, 105
stump cotton, 23
substitution of capital for labor, 313
suceptability of m.m. fibers to pilling, 388
suceptability of m.m. fibers, snagging, 388
suction scour m/c, 210
sufficiency of samples, 390
suint liquors, evaporated, 212
suint, 32, 208
supervisory costs, 308
surface abrasion of fabric, 388
symbols, 319, 324
synthetic fiber, history, 4
synthetic polymer, 20
synthetic staple & natural fiber blends, 5
systematic series, thick & thin places, 282
take-up rolls, 157
take-up speed, 45
tape condenser, 229, 230
tape drive, slippage, 458
tariffs and quotas, 8
tastes & ability to buy, 7
technical analyses, marketing, 277
technical analyses, problem solving, 277
technical fil, 4
technical service, 8
technology of production, 10
temperature distributions, 40
temperature, 341, 344
temperature, abs zero, 341
tenacity, 58
tensile tester, stiff system, 364
tensile testing of strands, 364, 368
tension control, 69, 217, 451
test fiber immaturity, 354
test laboratory, air-conditioned, 350
test rotor yarn, 379
testing by observation, 369
testing fiber during transit, 368
testing fil yarns, 351
testing fil yarns, dyed knitted sleeve, 362
testing for yarn defects, 290, 354, 355
testing linear density, weighing, 355
testing natural fibers, 279
testing of textile materials, 350
testing rotor yarn, untwisted, 359
testing single strands, 364, 366
testing staple yarns, 351, 360
testing textured yarns, 369
testing yarn bulk, water immersed, 370
testing yarn hairiness, 363
testing, 8, 276
testing, organization, 300
testing, supplier & supplied agreed, 280
tests influenced by r.h., 350
tex, 318
textile materials, 1, 18
textile products & fiber production, 18
texture, 10
textured yarn production, 12
textured yarn properties, 91, 368, 369
textured yarn structures, 362, 369, 383
textured yarn, 12
texturing calc, 339, 340
texturing, 4, 21, 56, 89, 90, 92–98, 110
texturing, air-jet, 13
texturing, co-extrusion, 112
texturing, cooling length, 94
texturing, disk twister, 99
texturing, false-twist & air-jet, 106
texturing, false-twist, 13, 98
texturing, fil handling, 94
texturing, friction twisting, 98
texturing, heater length, 94
texturing, knit-de-knit, bulk, 112
texturing, theoretical model, 90
texturing, threadline, 94, 95
texturing, tight spots, 95, 97
texturing, twist surges, 99
texturing, two-heater machines, 96
texturing, wear rates, 95
texturing, wild fil, 98
texturing, yarn attitude, 99
texturing, yarn quality, 98, 110
theoretical assignment, 308
theoretical CVs, fiber groups, 419
theoretical twist, 61
theory of blending capacity, 395
theory of doubling, 83
thin jets of polymer, 12
thin spots, twist, 177
throwing, 88
tight spots, 81
time domain, 356
titanium dioxide, 48, 95
TM/tpi relationship, 323
tongue-and-groove system, 418
toothed drafting, 122, 407, 412, 413, 457
top making, 217
top roll hardness, 156
top, 49
torque/twist curve, 383, 384
torsional stiffness, 68
Index
total variance, 281
tow and man-made staple, 13
tow knotting, 53
tow supplied to mills, 48
tow, 13, 15, 48, 49, 53
tow, drawing, 82
tow, stretch-break, 13
tow-to-top, 51, 220
tow-to-yarn, 51
tracer fiber, 457
tracing source of faulty bobbins, 255
transducer cost, 417
transducer time response, 417
transducer, 158, 159, 359, 364, 417
transfer costs of textile product, 6
transfer systems, automated, 7
transient can storage, 159
trans-oceanic shipping, 316
transport material, 184, 310
trash content, 281, 355
trash crushing, 127
trash in card flats, 284
trash reintroduced into fiber stream, 127
trash removal, 117
trash, 30, 117, 284, 286
traveler & balloon collapse, 443
traveler balance, 180, 181, 442, 443, 445
traveler change schedule, 180
traveler life, 180, 298, 442, 443
traveler mass vs yarn count, 443
traveler mass, see traveler weight
traveler moment balance, 442
traveler numbering, 181
traveler porpoising, 443
traveler scar, 443
traveler vs yarn tension, 443
traveler weight, 177, 181, 433, 440, 443
traveler, 177–183, 427, 433, 442–444
traveler, centroid, 443
traveler, fiber build-ups, 179, 283, 289
traveler, kinetic energy, 433
traveler/ring, friction, 443
traveling cleaner, 288
trend analysis, 359
trumpet condenser, 162
tube goemetry, 178, 180
tuft curves, 280
turbulent airflow, 106
twist & flow, 57
twist angle, 323
twist balance, 268
twist calcs, 335, 336
twist carried by moving yarn, 62
twist contraction, 257, 360
twist density, 57, 323, 324
twist direction, 57
twist evenness, 68
twist factor (α), 324
twist gear, 57
487
twist
twist
twist
twist
twist
twist
twist
twist
twist
twist
gradients, 177
in rotor spg, 193
insertion, 61
liveliness, 60, 68
migration, 68
multiple, 57–60, 206, 323, 324
projected by twister, 62
testing, 360, 446
to control fil, 61
triangle, 82, 174, 261, 262, 265, 314,
375, 378, 417
twist, 56, 61, 374
twist, staple yarn, 57
twist, thin spots, 69
twist/untwist method, 360
twisted self-twist yarns, 274
twisted yarn, compressive force, 374
twister speeds, 93
twist-gear, 169
twisting & doubling, 14, 220
twisting fil & frictional behavior, 383
twisting fil under heat, 383
twisting m/c, 250
two-for-one twisting, 63–67
two-heater machine, 93
typical process schedules, 15
ultra-fine fibers, 10
ultra-long errors in yarn, 391, 398
undercard waste, 287
undrawn polymers, 47
uneven breaking, 49
uneven build-up of finish on rolls, 47
uneven dye penetration, 114
uneven gripping of web, 163
uneven hardening, 73
uniform distribution, 346
uniformity index, 353
unit standard cell, 386
unopened (‘green’) bolls, 26
unraveled knitted fabric, 112
untextured fil, linear density, 368
untwisted yarn strength, 260
untwisting to zero twist, 359
unwinding, 243, 244, 451, 452
unwinding, balloon breakers, 244
unwinding, chaotic balloon, 244, 452
unwinding, end-breakage, 244
unwinding, hair plucked from adjacent yarns,
243
unwinding, tension controllers, 244
unwinding, yarn faults, 244
upper half mean length, 353
upper zone of balloon, 445
up-twisting, 65
USP, normalized comparisons, 295
Uster Classimat, 283
Uster evenness tester, 356
Uster hairiness values, 295
488
Index
Uster statistics, 295, 302
Uster statistics, percentile rankings (USP),
164, 295
utilities, 341
variable cost, 309–312
variable grip on fibers, 81
variable ratch setting, 81
variance, 69, 83, 281, 422, 424
variation period, 395
variation, skein-skein, 366
variations in hardness, 157
variations, annual, 393
vector alignment, 399
vegetable & mineral particles, 208
vegetable matter, 32, 34, 210
vinyl fibers, 38
virtual nip zone, 413
visco-elastic fibers & yarns, 365
visco-elastic polymer, 383, 384
visco-elastic rubber, 73
viscose rayon production, 40
viscose rayon solution, 4
viscosity, 39
viscous effects within polymer, 384
visual examination of fabrics, 372
visual examination of yarns, 361, 372
volume occupied, 12, 371, 343
volume of unit cell, 387
wage levels, elements of difference, 6
wage rates, 7, 316
wall-mounted hygrometers, 346
warehouse management, 391
warehouse stock, 391
warp beam defects, 289
warp wind, 169
waste disposal, scouring, 212
waste fiber, 126, 150, 151, 182, 286
waste, air currents, 152
waste, card, ss, 151, 152
waste, dust house, 151
waste, handled pneumatically, 151
waste, pepper trash, 152
waste, product flow calculations, 152
waste, reusable, 151, 152
wastewater disposal, 212
water vapor & steam, 341, 342
wavelength & frequency, 398
wax & sweat glands, 32
weak link, variable strengths, 365, 413
weaken fiber, 388
wealth flows, 7
wear of elements, 156, 417
wear resistant materials, 188
weathering, 26
weaving performance, 290
web doubling, 161, 163
weft wind, 177
weighpan feeder, 136, 123, 223
weight & mass, 317
wet & dry bulb temperatures, 344
wet spg, 38, 40
wet spg, extrudate, 40, 41
wet steam, 342
wet-spun fibers X-section, 40
wide band of technology, 219
widing & clearing, 235
wild fil, 61
willeyed blend, 220, 223
wind structure, 177
wind, precision, 242
winder productivity, 327, 328
winder, OE, 186, 197
winders, 15, 168, 234
winding & yarn dyeing, 257
winding complaints, prior processing, 278
winding costs, 304, 305
winding frames, ply, 251
winding heads/splicer, 246
winding m/c, 234, 235, 236, 241, 250, 254,
255, 315
winding m/c, air pollution, 246
winding m/c, automatic, 254
winding m/c, capital costs, 254
winding m/c, operator attention, 254
winding m/c, stable packages, 250
winding m/c, yarn joining, 254
winding m/c, wear, 250
winding performance, 327
winding tensions, 242
winding traverse mechanisms, 240, 242
winding traverse, pattern breaking, 240
winding traverse, ribboning, 240
winding, customer interface, 250
winding, damage resistance, 235
winding, defect removal, 245
winding, economics, 235
winding, faulty, 114
winding, hank, 236
winding, hard shoulders, 236
winding, invariable supply speed, 242
winding, nep, 250
winding, over-tension yarn damage, 242
winding, parallel w flanges, 236
winding, quality, 235
winding, reciprocating guide, 235, 242
winding, remove yarn faults, 235
winding, ribboning, 236
winding, skein, 236
winding, temporary yarn storage, 242
winding, tension control, 242, 245
winding, traverses per rotation, 236
winding, twisted yarn, 65
winding, withdrawal at will, 242
winding, yarn appearance, 250
winding, yarn clearing, 245
winding, yarn contraction, 246
Index
winding, yarn hairiness, 250
winding, yarn lag, 240
winding, yarn moisture content, 246
winding, yarn tension, 235, 236
winding. grooved roll, 242
wind-off speed, 65
windows, 349
winnowing, 127
wire flyer, 65
within-bale consistency, 391
within-clump variance, 390
wool & stretch-broken yarn, 50
wool & synthetic fibers, 3
wool classer, 33
wool cleaning, 32, 208
wool color, 32
wool crimp, 205
wool delivery in cool state, 210
wool devil, 221
wool dryer, 210
wool dyeability, 208
wool felting, 210
wool fiber blending, 205
wool fiber cleaning, 205, 211
wool fiber crimp, 32
wool fiber elongation, 32
wool fiber oiling, 210, 212
wool fiber strength, 32
wool fiber tips, 32
wool fibers, preparation, 206
wool grease by-products, 208
wool grease separation, 209
wool grease, 14, 208, 212
wool harvesting, 33
wool history, 2
wool kempts, 33
wool moisture content, 210
wool opener, 221
wool prices, 34
wool production, 33
wool rinsing and drying, 210
wool scouring, 208
wool shoddy, 286
wool tops, 34, 281
wool use, 33
wool willow, 221
wool yarn, crease recovery, 381
wool, 20, 31, 33, 34, 205, 210, 212, 213, 382
wool, blended, 213
wool, carbonizing, 210
wool, core-boring, 279
wool, cotted, stained or muddied, 34
wool, crease-shedding properties, 3
wool, cuticle & cortex, 33
wool, deburred, 213
wool, fiber diameter, 280
wool, fiber length, 33
wool, fiber strength loss, 210
wool, fleeces classification, 34
489
wool, foreign matter, 208
wool, grease, 213, 221, 284
wool, greasy, 32, 208, 213
wool, history, 3
wool, locks, 221
wool, market size, 3
wool, matchings, 34
wool, medulated fibers, 33
wool, moisture absorption, 3
wool, murrain, 33
wool, natural finish removal, 381
wool, pieces, 221
wool, prickle, 32
wool, removal of foreign materials, 32
wool, sampling procedure, 279
wool, scour, 34, 212, 213
wool, shorn or pulled fiber, 33, 208
wool, shorn or sheared, 33
wool, slipes, 33
wool, suint, 284
wool, tender, 31
wool, variability, 34, 221
wool, vegetable matter, 284
wool, washed & rinsed, 208
wool, wax & suint, 32
wool/m.m.fiber blend, 205, 216
woolen blending, 220, 222
woolen card delivery, 229
woolen card set, 224, 228
woolen carder, 226
woolen carding, feed, 220
woolen carding, initial, 223
woolen spg m/c, damage, 212
woolen spg, autocount system, 228
woolen spg, balloon control rings, 231
woolen spg, blending procedures, 220
woolen spg, blending, 211, 223, 228
woolen spg, carder (finisher card), 225
woolen spg, chute feed, 224
woolen spg, clean & sterilize fibers, 211
woolen spg, cleaning system, 211
woolen spg, cleaning, 221
woolen spg, collapsed balloons, 231
woolen spg, color differences, 220
woolen spg, computer control, 224
woolen spg, condensed slubbing, 221
woolen spg, control signals, 228
woolen spg, cross-lapper, 225
woolen spg, crush roll set, 225
woolen spg, dancing roller, 224
woolen spg, decomposition of raw material,
221
woolen spg, doubling (plying), 231
woolen spg, draw (pause), 221
woolen spg, equipment groupings, 224
woolen spg, fearnought, 221, 222, 223
woolen spg, feed rolls, 225
woolen spg, fiber baling, 223
woolen spg, fiber degradation, 228
490
Index
woolen spg, fiber electrification, 211
woolen spg, fiber handling, 222, 223
woolen spg, fiber loss, 228
woolen spg, fiber oiling, 211, 222, 228
woolen spg, fiber pollution, 224
woolen spg, flexible wire, 226
woolen spg, fractionation, pinned rolls, 224
woolen spg, gamma ray thickness sensor, 224
woolen spg, garnet clothing, 225
woolen spg, hopper feeders, 222, 223
woolen spg, human monitoring, 224
woolen spg, lattice feed, 225
woolen spg, m.m. fiber, 223
woolen spg, manual blending, 222
woolen spg, mass flow control, 224
woolen spg, mechanical layering, 222
woolen spg, mechanical mixing, 222
woolen spg, necessity for blending, 220
woolen spg, oiling system, 223
woolen spg, opening, 221, 228
woolen spg, periodic dumping, 223
woolen spg, picker opener, 221
woolen spg, processes variable, 220
woolen spg, r.h., 211
woolen spg, reworking, 222
woolen spg, rotary spreader, 222
woolen spg, safety, 226
woolen spg, sampling & testing, 223
woolen spg, Scotch feed, 225
woolen spg, scribbler (breaker card), 225
woolen spg, sensors, 224
woolen spg, settings, 228
woolen spg, short process, 221
woolen spg, sliver, 229
woolen spg, slubbing condenser, 224
woolen spg, slubbings, 225, 229
woolen spg, spike size, 221
woolen spg, spin limit, 228
woolen spg, stock records, 222
woolen spg, stock-dye fibers, 222
woolen spg, swift loadings, 228
woolen spg, swift/worker/stripper, 225
woolen spg, synchronized production, 223
woolen spg, tape condenser, 225, 226, 229
woolen spg, teaser, 221
woolen spg, thickness sensor, 224
woolen spg, threads, 221
woolen spg, tooth size, 221
woolen spg, tumbling, 223
woolen spg, variety cleaning systems, 211
woolen spg, varying fiber cleanliness, 223
woolen spg, waste fibers, 220
woolen spg, web error, 228
woolen spg, workers & strippers, 225
woolen system, 15, 205, 219, 220
woolen yarns, 205, 231
work practices in mill, 291
worked examples, 329
worker exploitation, 6
workspace r.h., 282, 346
worsted hank, 318
worsted processing system variations, 213
worsted S-on-S or Z-on-Z ply yarns, 268
worsted spg, 15, 218
worsted spg, automatic winding, 219
worsted spg, bobbin length, 219
worsted spg, bobbin movement, 219
worsted spg, constant balloon length, 219
worsted spg, electronic clearing, 219
worsted spg, large packages, 219
worsted spg, m.m. fibers, 216
worsted spg, ratch settings, 216
worsted spg, ring size, 219
worsted spg, splicing, 219
worsted spg, tension, 219
worsted spg, twist levels, 216
worsted system, 205, 213, 216, 219
worsted, plied warp yarns, 268
wound packages, series of specimens, 287
woven fabric simulation, 362
wrap friction, 75
wrap spg bouclé yarns, 270
wrap spg false twist in staple, 270
wrap spg fil wraps core, 270
wrap spg fil yarn via hollow spindle, 270
wrap spg productivity, 270
wrap spg rotating hook, 270
wrap spg, 66
wrappings, 22
wrap-spun yarns, 270
wrap-spun yarns, core twist, 270
wrinkle resistance, 2
yard board, 355, 356, 361
yarn & strand numbering, 317
yarn amount to be tested, 290
yarn appearance, 278, 282, 361
yarn balloon base, energy, 445
yarn balloon control, 65
yarn balloon energy balances, 427
yarn balloon mechanics, 427
yarn balloon, 70, 177, 188
yarn balloon, air-drag, 432, 437
yarn balloon, air-flow, moving parts, 438
yarn balloon, ambiguous tension, 443
yarn balloon, balloon collapse, 435
yarn balloon, balloon shape, 435
yarn balloon, base, 428
yarn balloon, central zone, 427, 436, 439
yarn balloon, centroid, 429, 448
yarn balloon, chain of elements, 448, 449
yarn balloon, changes in energy level, 434
yarn balloon, chase motion, 432
yarn balloon, collapse, 434
yarn balloon, conservation of energy, 431
yarn balloon, crown, 450
yarn balloon, cyclical traveler speeds, 440
yarn balloon, drag coefficients, 438
Index
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
balloon, elements tilt & twist, 437
balloon, end conditions, 429, 450
balloon, energy balance, 431
balloon, energy equilibrium state, 448
balloon, energy to inflate balloon, 444
balloon, energy, 430, 431, 433, 434, 448
balloon, equilibrium, 450
balloon, experimental data, 440, 441
balloon, force analysis, 429, 432, 436,
437, 440, 444
balloon, force analysis, 449, 450
balloon, forces, 427
balloon, friction, 429, 435, 445
balloon, frictional restraints, 436
balloon, geometry changes, 445
balloon, gradients, 437
balloon, hairy comet’s tail, 438
balloon, inertial frame, 440
balloon, instabilities, 434
balloon, insufficient energy, 434
balloon, kinetic energy, 431
balloon, lay point, 431
balloon, loci of segments, 447
balloon, lower zone, 427
balloon, multi-noded, 435
balloon, non repetitive, 440
balloon, offset pigtail guide, 445
balloon, periodic collapses, 448
balloon, pigtail guide, 445
balloon, plan view, 440
balloon, protruding hairs, 437
balloon, radius of gyration, 434
balloon, radius of locus, 450
balloon, radius of yarn curvature, 450
balloon, ring lubrication, 450
balloon, rotating plane, 428
balloon, rotating reference plane, 437
balloon, rotational speed, 451
balloon, separator plates, 437
balloon, shape changes, 447, 448
balloon, simplified, 428
balloon, single-noded, 434
balloon, size & shape, 434
balloon, speed stability, 447
balloon, strain energy, 431
balloon, subsidiary oscillations, 434
balloon, surface swept, 427
balloon, temporary collapses, 448
balloon, tension & twist of yarn, 445
balloon, tension at wind point, 434
balloon, tension gradients, 435, 449
balloon, tension variations, 445
balloon, tension, 435, 445
balloon, torque supplied, 431
balloon, torque, 445
balloon, torsional strain energy, 433
balloon, traveler centroid, 442
balloon, traveler instability, 434, 448
balloon, traveler mass, 434, 450
yarn
yarn
yarn
yarn
491
balloon, traveler, 442
balloon, twist gradients, 446
balloon, upper zone, 427
balloon, vector analysis, 427, 431, 439,
448
yarn balloon, vertex, 428
yarn balloon, winding tension, 444
yarn balloon, wind-off point, 451
yarn balloon, yarn tension, 439
yarn blackboard, 361
yarn blend, 277
yarn bobbins, 175
yarn bulk, 89, 384, 386
yarn bulk, geometry fil helices, 386
yarn cheese, 65
yarn clearing vs winding efficiency, 255
yarn clearing, 234, 235, 245, 246, 247, 287
yarn clearing, automatic splicer, 245
yarn clearing, capacitive sensors, 246
yarn clearing, defect sources, 245
yarn clearing, nub plates, 245
yarn clearing, optical devices, 246
yarn clearing, patrolling piecer, 245
yarn clearing, performance, 247
yarn clearing, prescribed limits, 245, 246
yarn clearing, residual faults, 255
yarn clearing, settings, 246
yarn clearing, size of defect removed, 287
yarn clearing, wind at high tension, 245
yarn color, 114
yarn conditioning, steaming, 246, 257
yarn cone, 65
yarn contraction, 257
yarn cooling, 384
yarn cost, end-breakage, 314
yarn cost, spg a major portion, 314
yarn count, 314, 318
yarn cross-sections, ring and rotor, 381
yarn CV, 254, 295, 458
yarn damage, 180
yarn defect classification, 290
yarn defects, 173, 227, 243, 254, 277, 281–
283, 286, 288, 291, 348
yarn degradation, 153, 300
yarn diameter, 374
yarn dyeing performance, 258, 278
yarn economics, 301
yarn elongation, 424
yarn error, test length, 294
yarn errors, fiber flow, 291
yarn errors, mechanical, 291
yarn evenness, 277, 291
yarn evenness, OE, 455
yarn failure, fiber breakage, 58
yarn fineness, 317
yarn from rotating package, 65
yarn grade, 361
yarn hairiness, 19, 176, 178, 181, 263, 277, 281,
288, 295, 375, 377, 380, 381, 417, 448
492
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
yarn
Index
hairiness, m/c type, 381
hairiness, reduction, 262
hairiness, within-bobbin variation, 295
hand, 384
heated above Tg, 385
hysteresis, 383, 384
joining, 220, 245, 277
length on packages delivered, 284
length/sq yard fabric, 290
manufacturing problems, 254
moisture absorption, 257
number, 318
numbering calcs, 330, 331
or tow take-up rate, 45
package build, 65, 177
package density, 181
package diameter, 180
package inside the balloon, 65
package types, 234
processability, 277
properties, OE, 190, 194, 195, 260
quality & operator assignment, 306
quality, 17, 316
relaxation, 381, 385
removal from rotor groove, 314
removal, OE, 186
slub, 227, 282
sources, 302
splice strength, 246, 250
splice, dynamometer, 250
splice, splice efficiency, 250
splice, stiffness, 250
splice, tails, 248
splice, thick spot, 248
splicer, hand-held, 246
splicer, multiple, 246
splicer, settings, 248
splicing, 56, 235, 246–249
splicing, component wear, 249
splicing, end conditioning, 247–249
splicing, joint appearance, 248
splicing, lint accumlation, 249
splicing, rubber elements, 249
splicing, temporary false twist, 248
spun from card, 205
strength regularity, 296
strength, 163, 277, 281
strength, fiber force components, 373
structure, 89, 377
tails, packages, 258
tenacity CVs, 296
tension control, 69
tension gradients, 435
tension, 65, 177, 178, 180–183, 218,
243, 314
testing, 290, 365
thin spots, 306
transfers, 243
twist liveliness, 277, 381
yarn twist, 19
yarn twist, exit from twist triangle, 446
yarn twisting, OE, 186
yarn untwist, 384, 385
yarn value, reduction, 128
yarn variance, 365
yarn weak spots, 289
yarn winding point, ring rail motion, 444
yarn, air-jet interlacing, 275
yarn, air-jet, see air-jet yarn
yarn, appearance, 89
yarn, blended (dissimilar fibers), 155
yarn, blended, 13, 155
yarn, bulk, 12, 61, 92, 93, 371, 385
yarn, co-extruded, 109
yarn, co-mingled, 109
yarn, core, 109
yarn, core/sheath, 112, 260
yarn, creel breaks, 173
yarn, cross-sectional shape, 321
yarn, CV, 188
yarn, elastomeric, 112
yarn, elongation, 89, 112
yarn, false-twisted, 61, 177
yarn, fault removal, 234
yarn, fibrillated, 109
yarn, fine linen, 232
yarn, flat fil, 109
yarn, frictional behavior, 381
yarn, hand, 89
yarn, hand, OE, 195
yarn, heat setting, 89
yarn, heat-set when twisted, 384
yarn, high-bulk, 50
yarn, interfiber friction, 89
yarn, linear density, 177
yarn, low-bulk, high-stretch, 91
yarn, ls, folding, 220
yarn, ls, spun from sliver, 220
yarn, mock ply, 261
yarn, open-end, 185
yarn, over-conditioning, 257
yarn, ply, 14
yarn, self twist, 261, 271, 273
yarn, self twist, piecing, 273
yarn, Selfil, 275
yarn, short-term errors, 297
yarn, slit, 109
yarn, slub, 173
yarn, soft-wound packages, 95
yarn, squashed, 321
yarn, staple/fil, 260
yarn, STT, character, 274
yarn, STT, composite ply yarns, 275
yarn, STT, modified, air-jet texturing, 275
yarn, STT, real twist addition, 274
yarn, textured, 61
yarn, theoretical diameter, 322
yarn, thermoplastic, 89
Index
yarn, thick & thin spots, 173
yarn, twisted self-twist (STT), 274
yarn, twist-liveliness, 277
yarn, winds at differing diameters, 17
yarn, woolen, 220
yarn, worsted, 205
yarn, wrap-spun, 270
yarn-making technologies, modern, 11
yarns of complex structure, 260
yarns, cable, 61
yarns, carpet, 15
yarns, core/sheath, 109
yarns, fancy, 219
yarns, industrial, 16
yarns, ls, heavy, 206
yarns, modified twist, 261
yarns, plied acrylic, hand knitting, 253
yarns, plied aramid, ropes etc., 253
yarns, plied, costs, 253
yarns, silk, 113–114
yarns, viscose rayon, 381
yarns, world market, 9
yarns, worsted, 219
yellowness (+b), 355, 404, 406
493