Uploaded by sedat kumartaşlı

jaffe2020 (1)

Polyester fibers
8
Michel Jaffe1, Anthony J. Easts2 and Xianhong Feng3
1
New Jersey Innovation Institute, University Heights, Newark, NJ, United States,
2
Consultant, Madison, NJ, United States, 3Beckton Dickinson and Company, Franklin
Lakes, NJ, United States
Abstract
Polyester fiber, specifically poly(ethylene terephthalate) (PET) fiber, is the largest volume
synthetic fiber produced worldwide. The total volume produced in 2016 exceeded 50 million tons with a rate of growth far greater than any other fiber, natural or synthetic. Low
cost, convenient processability, ease of blending with cotton and other natural fibers, convenient recyclability, and excellent and tailorable performance are the reasons for the
dominating success of PET fiber.
The excellent performance of polyester fiber over a wide range of end uses results from
the ability to accurately control fiber morphology (distribution and connectivity of crystalline and noncrystalline load-bearing units) allowing the balance of thermal and dimensional stability, transport, and mechanical properties to be precisely controlled. All these
parameters are conveniently and accurately monitored by thermal analysis techniques. It
is the purpose of this chapter to provide the reader with an overview of the applications of
thermal analysis toward polyester fiber characterization, including the impact of processing on performance and the utility of thermal analysis toward understanding the materials
science of PET fibers.
8.1
Introduction
Polyester fiber, specifically poly(ethylene terephthalate) (PET) fiber, is the largest
volume synthetic fiber produced worldwide. The total volume produced in 2016
exceeds 50 million tons with a rate of growth far greater than any other fiber, natural or synthetic. The distribution of synthetic fiber production by chemistry is
shown in Fig. 8.1 (Industrievereingigung Chemiefase, 2018).
If one assumes the total production is a single 5 dpf (B20 μm diameter) filament, the total length would be measured in light years (B1016 m) or the distance
equivalent of leaving our solar system. While other polyesters are commercially
produced in fiber form—polyethylene naphthalate (PEN), poly(butylene terephthalate) (PBT), poly(propylene terephthalate) (PPT), poly(lactic acid), and thermotropic
polyester (liquid crystalline polymer—see chapter 19, Thermal analysis of liquid
crystalline polymers)—these are of insignificant volume compared to PET; hence,
Thermal Analysis of Textiles and Fibers. DOI: https://doi.org/10.1016/B978-0-08-100572-9.00008-2
© 2020 Elsevier Ltd. All rights reserved.
134
Thermal Analysis of Textiles and Fibers
Figure 8.1 Production of Polyester Fiber from 1975 2017 (in 1000 metric tons)
(Industrievereingigung Chemiefase, 2018).
this chapter will focus primarily on PET. Recently, poly(ethylene furanoate)—PEF
(the polyester of ethylene glycol and 2,5 furandicarboxylic acid), has been discussed as a commercial competitor to PET, but the emphasis is more on food packaging than fiber.
The reasons for the dominating success of PET fiber are as follows:
G
G
G
G
G
Low cost
Convenient processability
Ease of blending with cotton and other natural fibers
Convenient recyclability
Excellent and tailorable performance
The origin of low-cost polyester fiber lies in the efficient conversion of mixed
xylenes to terephthalic acid; for details, see Modern Polyesters: chemistry and technology of polyesters and polyesters (Scheirs and Long, 2005). Cost and desirability
is also positively impacted by the 280 C melting temperature of high-molecular
weight PET, which allows the use of commercial heating fluids for processing, and
the 75 C glass transition enables the morphology and molecular orientation introduced during processing to be stable at room temperature through washing temperatures. The excellent performance of polyester fiber over a wide range of end uses
results from the ability to accurately control fiber morphology (distribution and connectivity of crystalline and noncrystalline load-bearing units), allowing the balance
of thermal and dimensional stability, and transport and mechanical properties to be
precisely controlled. All these parameters are conveniently and accurately monitored by thermal analysis techniques (Jaffe et al., 1981, 1997). By the end of the
20th century, PET fiber manufacture had shifted from the United States and Europe
to Asia, with China and India dominating PET fiber manufacture, reducing the
study of PET in the west but catalyzing research in the east. The quest for increased
product sustainability has fostered the commercial development of PPT in the
United States (see SCI, 1999; Scheirs and Long, 2005) and, more recently, PEF in
the United States and Europe. Polyester fiber technology and performance have
been reviewed in many publications (Eichhorn et al., 2009) and the reader is
Polyester fibers
135
directed to these publications for additional detail. It is the purpose of this chapter
to provide the reader with an overview of the applications of thermal analysis
toward polyester fiber characterization, including the impact of processing on performance and the utility of thermal analysis toward understanding the materials science of PET fibers.
8.2
Poly(ethylene terephthalate) history
The development of PET fiber, as with all synthetic thermoplastic fibers, originates
with the research into aliphatic condensation polymers led by W. H. Carothers of
DuPont in the 1930s (McIntyre, 2005). Much improved fiber performance was
achieved in the early 1940s by the British team of J. Rex Whinfield and J. T. Dickson
(Cheremisinoff, 1989), who focused their work on the aromaticaliphatic polyester of
terephthalic acid and ethylene glycol. Commercialization of PET was rapid after
World War II with the introduction of Terylene in Great Britain by ICI and the introduction of Dacron in the United States by DuPont, and PET fiber successfully entered
the textile market as both filament yarn and staple and in the industrial market as
a rubber reinforcement filament yarn, primarily for use in the sidewalls of radial
passenger-car tires. Key performance advantages were wash-and-wear characteristics,
blendability with cotton in textiles, and high modulus, coupled with excellent modulus
retention, in industrial applications. Continued progress in the efficient, low-cost
production of terephthalic acid ensured the dominance of PET as the fiber of choice in
most fiber applications.
8.3
Poly(ethylene terephthalate) polymerization
PET is the condensation product of terephthalic acid and ethylene glycol. The key
to successful PET polymerization is monomer purity and the absence of moisture in
the reaction vessel. PET polymerization has often been reviewed (Burghardt and
Vom, 1974) and the reader is referred to the many journal articles and patents dealing with all aspects of polyester fiber production (see, e.g., Nichols et al., 1996;
Paszun and Spychaj, 1997). The first stage of PET polymerization is, in essence,
the production of bis(hydroxyethyl)terephthalate (BHET). In the direct esterification
of terephthalic acid, the reaction
HOOC 2 C6 H4 2 COOH 1 2HOCH2 CH2 OH ! HOCH2 CH2 OCO
2 C6 H4 2 COOCH2 CH2 OH 1 H2 O
(8.1)
results in a mixture of low amounts of free BHET with a variety of PET oligomers. Water removal, usually under high vacuum conditions, is critical to the
ultimate achievement of high molecular weights. The reaction catalysts for the
136
Thermal Analysis of Textiles and Fibers
ester interchange reaction have been the subject of intense research for many
years and many catalyst compositions may be found in the patent literature
(Wilfong, 1961; French Patent, 1959; Easley, 1959). The introduction of ester
interchange catalysts requires the killing of these catalysts later in the polymerization sequence as they are equally effective as depolymerization catalysts. The
next step in PET polymerization is melt condensation to high molecular weight.
In this reaction step an ester interchange reaction occurs between two molecules
of BHET to split off a molecule of glycol, building polymer molecular weight.
The reaction must be catalyzed, and antimony trioxide, Sb2O3, is almost universally the moiety of choice. Typical melt-polymerization temperatures are at or
above 285 C and viscosities are in the order of 3000 P, making uniform stirring
and the imparting of uniform shear history, across the polymerization mixture,
difficult to achieve. There is a prolific patent literature describing variations and
improvements to PET polymerization. For reviews, see the work of Eichhorn
et al. (2009) and East (2004).
After achieving target molecular weight, the polymer can be pelletized for subsequent melt spinning (batch processing) or fed directly into a spinning machine and
converted to fiber (continuous processing)—if the spun fiber is fed directly to a
draw frame, the process is known as continuous polymerization (CP) spin-draw.
The molecular weight of PET pellets can be further increased through solid-state
polymerization. In this process, thoroughly dried PET chip is first crystallized at
about 160 C (to prevent the as-polymerized chip from sticking together) and then
heated just below the melting point under high vacuum and extreme dryness to
advance the molecular weight upwards to values of inherent viscosity (IV) of 95
(textile grade chip has an IV of about 0.65) (Callender, 1985; Sorenson et al.,
2001). The effects of the thermal-history process of PET chip and fiber have been
extensively studied and are conveniently monitored by thermal analysis techniques.
Jaffe et al. (1997) have reviewed the thermal behavior of PET and described the
expected response of PET to process history in detail.
8.4
Characterization of poly(ethylene terephthalate) chip
PET chip or representative samples of CP spin-draw polymer are conveniently characterized by molecular weight, cleanliness, and thermal behavior. Molecular weight
is characterized by the polymer IV, usually in halogenated solvents, and while the
IV is related to molecular weight by the MarkHouwink equation, this is seldom
done in practice.
Polymer cleanliness is measured microscopically (optical techniques, polarized light microscopy) and is often expressed in subjective units such as the
average number of black specks or the number of gels per gram of polymer.
Cleanliness of the polymer is critical, a particle or gel of only a few microns
in diameter can be responsible for a catastrophic spinline interruption. Thermal
parameters are conveniently monitored by differential scanning calorimetry
Polyester fibers
137
(DSC), allowing a quick assessment of diethylen eglycol (DEG) content,
crystallinity, etc. (Jaffe et al., 1997; Eichhorn et al., 2009).
8.5
Poly(ethylene terephthalate) fiber processing
The melt spinning of PET has been extensively treated in the patent literature,
somewhat less so in the open literature (Davis and Hill, 1982), although the chapters by East (2004), Bessey and Jaffe (Ward et al., 2000), and Reese (2003) are
good introductions to the process. Here, we concentrate on how changes in the key
process variables of spinline stress and temperature profile affect the morphology
developed in fiber spinning and drawing, and in turn, how the morphology affects
the resulting performance of the yarn as manifested in thermal analysis. Process
issues have been discussed in Chapter 2, Fiber processstructureproperty relationships, and that discussion is directly relevant to polyester-fiber production. The
key process and performance parameters will be linked to the thermal analysis
methods best utilized to elucidate the level and origin of the parameter of interest.
The extraordinarily high volume of PET fiber produced in a given manufacturing
plant, coupled with the huge length of fiber per unit mass, places increased emphasis on the statistics of any measurement made on fiber samples. In reality a coefficient of variation of 6 10% or less for key parameters (diameter, tensile
properties, shrinkage) is required for the yarn to be commercially acceptable. For
example, changes in noncrystalline molecular chain orientation (as manifested in
fiber shrinkage at Tg), which is directly related to dye uptake and variations as low
as 6 5%, may result in unacceptable lack of color uniformity in dyed fabrics. The
frequency of variation in processing is also critical; high-frequency changes that
may be averaged over a critical yarn length, are, in general, more acceptable than
smaller variation along or between specific filaments, which occurs at the lower
frequency.
The materials science of synthetic-fiber spinning discussed in Chapter 2, Fiber
processstructureproperty relationships, defines how morphology develops during spinning and drawing and is directly applicable to PET fiber. The remainder
of this chapter will focus on the specifics of the thermal analysis of PET. Specific
process details tend to be less discussed in the open literature and the reader is
thus directed to the patents of Celanese, DuPont, Fiber Industries, and Allied
Chemical Corporation (none of these companies currently exist) and the more
recent patents of Reliance, in India, and, for example, the Jiangsu Sanfangxiang
Group, in China.
Much of PET yarn production is converted to “staple” fiber and the demands of
staple fiber are different to those of filament yarns. Staple fiber is a continuous filament cut into short lengths of B30100 1 mm. Staple fibers are discontinuous
and are crimped and chopped to the desired staple-fiber length to effectively blend
with cotton, wool, or other natural fibers. The raw polyester fibers are melt-spun
through many hundred spinneret holes and collected in large drums or cans; fiber
138
Thermal Analysis of Textiles and Fibers
bundles from many cans are combined into thick bundles of fibers called a “tow,”
which is often of several million decitex. These thick bundles of fibers are then
drawn on a massively constructed draw frame, heat-set in a steam-heated hot box,
and then usually crimped using a “stuffer box” method. The bulked tow is finally
cut to the desired staple length and compressed into bales.
After spinning and drawing are completed, polyester yarns become components
in a variety of end-use products with textiles (fabrics) and rubber reinforcements
(tire cord) comprising the highest volume. Every end use involves subjecting the
fiber to additional temperature and stress history, impacting the observed thermal
response of the fibers being tested. This history is manifested in DSC, thermo
mechanical analysis (TMA), and differential thermo mechanical analysis (DTMA)
and examples are given in the discussions below.
8.6
Physical properties of poly(ethylene terephthalate)
PET is a semicrystalline polymer and its physical parameters have been repeatedly
determined over many years. Table 8.1 is a summary of typically accepted values
(Kitano et al., 1995).
8.7
Other polyesters
Other aromatic, aliphatic polyesters of commercial import include PEN, PPT, and
PBT. These are produced in far smaller volumes than PET and are often focused
toward specific markets: PEN for cordage, PPT for carpets, and PBT for stretched
fabrics. PEF fiber (patents go back over 60 years) is similar in performance to PET
but is not yet commercially produced. Initial applications for PEF are focused on
food packaging, but if commercial introduction is successful, the conversion of PEF
to fiber products is likely.
Table 8.1 Physical Parameters of PET.
Crystal habit
Cell parameters
Cell density
Tm (DSC)
ΔHf
Tg (solid chip)
Tg (drawn fiber)
Specific gravity
Triclinic: one polymer chain per unit cell
a 5 0.444 nm; b 5 0.591 nm; c 5 1.067 nm, α 5 100 degrees;
β 5 117 degrees; γ 5 112 degrees
1.52 g/cm3
260 C265 C
140 J/g; 33.5 cal/g
79 C (DSC)
120 C (dynamic loss)
1.33 (amorphous, undrawn), 1.39 (crystalline drawn fiber)
Polyester fibers
8.8
139
Thermal analysis (TA) of polyester fibers
The Thermal analysis (TA) literature associated with PET fiber is extensive and has
been thoroughly reviewed by Jaffe (Jaffe et al., 1981, 1997), and others (Eichhorn
et al., 2009) from the inception of modern thermal analysis in the 1960s to about
2010. The physical parameters of PET (TgB79 C, TmB175 C) and their molecular
relaxation times and crystallization rates enable convenient control of fiber morphology and allow the production of fibers for a broad range of applications.
Modern PET is usually spun at very high speeds of $ 5000 m/min; so, examples of
thermal analysis results that show the impact of spinning speed (spinline stress) will
be emphasized. Fig. 8.2 shows typical PET DSC behavior with Tg, crystallization
of heating (cold crystallization), melting, crystallization upon cooling, and, if heated
above B325 C, decomposition.
All phenomena associated with the physical changes, which are known to occur
in semicrystalline polymers, are noted in DSC studies of PET. Changes in process
history manifest in the location, size, and shape or the resulting DSC traces. For
example, changes in the melting of as-spun PET, as a function of spinning stress
(or spinning speed, molecular weight (MW), etc.), is shown in the classic work of
Heuvel and Huisman (1978) in Fig. 8.3. Fig. 8.3 clearly shows that as the spinning
speed (spinning stress), applied to the spinning PET fiber, is increased, the temperature of cold crystallization goes down toward Tg and the spinning PET fiber
decreases in size, while the crystallinity of the spun yarn increases.
Fig. 8.4 shows the premelting observed in PET fibers as a function of yarn
thermal-history (annealing temperature). The premelt endotherm represents the
Figure 8.2 DSC of as-spun PET fiber (A) is the first heating of a typical as-spun fiber, (B)
is the first cooling after melting, (C) is the second heating (Ms. Cindy Lee, Ms. Berta
Marani, Dr. Michael Jaffe, New Jersey Innovation Institute, unpublished results). PET, Poly
(ethylene terephthalate).
Figure 8.3 DSC of PET spun yarn as a function of spinning speed. PET, Poly(ethylene
terephthalate).
Source: From Heuval and Huisman (1978), J. Appl. Polym. Sci. 22(8), 22292243, reprinted
with permission from John Wiley & Sons, Ltd.
Figure 8.4 “Premelt” endotherm of commercial PET fiber as a function of thermal history.
PET, Poly(ethylene terephthalate).
Source: From Berndt and Bossman (1976), Polymer 17(3), 241245, reprinted with
permission of Butterworth Heinemann, Ltd.
Polyester fibers
141
melting of crystals formed during a heat-treatment step of the fiber or fabric; these
crystals are stable up to the temperature of formation and provide useful diagnosticprocess information. Such studies, pioneered by the late Edith Turi on nylon fibers,
give useful insight into the time and temperature history of polyester fibers.
Monitoring the dimensional stability (TMA) and mechanical properties of polyester fiber yields complementary information of DSC and provides insight into the
stress history of the fibers under investigation while simultaneously providing data
of end-use significance (shrinkage at a given temperature). Fig. 8.5 shows a diagrammatic representation of the dimensional change, occurring as an oriented and
crystallized PET fiber, which is allowed to shrink in a close-to-zero tension fiber
experiment. Process history affecting behavior at Tg, Tm, and processing temperatures manifests in the resulting plots of dimensional change versus temperature.
TMA can be utilized to measure the coefficient of linear expansion (CTE) in the
direction of the long axis of the fiber. It has been established that as the molecular
orientation of a fiber increases, the CTE decreases from the isotropic value of CVE/
3 to the change in the crystal unit cellchain dimension change in the fiber direction. As illustrated in Fig. 8.6, when measured in a region of temperature below Tg,
a useful way is provided to monitor average molecular orientation in the fiber,
which is especially useful if the fiber cross-section is not circular.
Examination of the dimensional stability changes in as-spun PET yarns clearly
shows the structural changes occurring as the line speed (spinning stress) is
increased. Referring again to the classic work of Heuval and Huisman (1978), it is
shown in Fig. 8.7, showing first the increase of molecular orientation with increasing spinning speed, accompanied by an increase in fiber initial modulus and, as
higher spinning speed, the introduction of crystallinity in the spinline, shown by the
shifting of shrinkage from Tg to Tm. Comparing Fig. 8.7 to Fig. 8.3 illustrates the
Figure 8.5 Diagrammatic representation of the dimension changes in processed PET fiber as
a function of process history (Jaffe, Celanese Research Company, unpublished results). PET,
Poly(ethylene terephthalate).
142
Thermal Analysis of Textiles and Fibers
Figure 8.6 Coefficient of linear thermal expansion of a PET fiber as a function of average
molecular orientation (Jaffe, Celanese Research Company, unpublished results). PET, Poly
(ethylene terephthalate).
Figure 8.7 Shrinkage of as-spun PET yarn as a function of spinning speed (Heuval and
Huisman, 1978). PET, Poly(ethylene terephthalate).
Source: Reprinted with permission of John Wiley & Sons.
Polyester fibers
143
detailed process understanding that can be obtained when the thermal response of
variously processed fibers is examined by a variety of complementary techniques.
Dynamic mechanical analysis of polyester fibers yields useful end-use characterization of polyester fibers by providing information to fiber modulus as a function
of temperature as well as a measure of Tg and a measure of the work loss exhibited
by the fiber during deformation. Fig. 8.8 represents the typical DMA data on PET
Figure 8.8 DMA of a drawn PET fiber as a function of the draw ratio (Miller and
Murayama, 1984). PET, Poly(ethylene terephthalate).
Source: Reprinted with permission of John Wiley & Sons, Ltd.
144
Thermal Analysis of Textiles and Fibers
fibers as a function of the applied draw ratio, showing the increase in initial fiber
modulus and the change in shape of tan δ.
The examples shown before illustrate how the thermal analysis of PET fiber
reflects the details of the fiber morphology and can be related to process history
and end-use performance. An excellent illustration on how this data may be used is
illustrated in the early work of the Valk et al. (1980) shown in Fig. 8.9, which
shows the relationship of fabric dye uptake to the annealing stress and temperature
applied to the fabric during heat setting (annealing under tension). One could substitute fabric shrinkage at a given temperature or the amorphous molecular orientation function for the dye uptake axis with similar results. If the fabric was made of
a polyester fiber with a different process history, the details of the plot would
change to reflect these differences.
The examples shown illustrate the power of thermal analysis in the characterization of PET fibers. Inherent in these examples are the several ways in which the
similar PET fiber thermal data can be utilized by the fiber scientist:
1. Finger printing: Does the fiber, produced today, match the thermal characteristics of fiber
produced yesterday and does it also meet the product specifications?
2. Quality control: Do the end-use properties, such as shrinkage at a given temperature
(TMA), coefficient of thermal expansion along the fiber axis (TMA), initial modulus or
hysteresis as a function of stretch/relax cycles (DMA), or overall structural stability as
measured by crystallinity (DSC) meet product specifications?
Figure 8.9 Relationship of fabric dye uptake to fabric heat-setting process history.
Source: From Valk et al. (1980), Textile Research Journal, 50(1), 4654, reprinted with
permission of Deutcher Verlag Gmbh.
Polyester fibers
145
3. Process development/diagnostic: A highly efficient tool to assist new product development and/or a diagnostic for what is causing a given process to yield off-specification
materials.
Similar techniques and experimental approaches can be applied to other
aromaticaliphatic polyesters of fiber importance including PEN, PPT, PBT, and
perhaps in the future, PEF.
8.9
Polyester fiber thermal analysis in the 21st century
As polyester fiber products became mature, studies of the thermal behavior of PET
and other polyester fibers became less frequent and tended to focus on specialized
rather than general aspects of polyester fibers.
In a major study of PET industrial yarns, Liu et al. (2016) compared the morphology of HMLS (high modulus low shrinkage) yarn to that of HMLE (high modulus low elongation). HMLS is produced from a highly oriented precursor yarn,
while HMLE begins with low orientation spun yarn; both are drawn to maximum
draw ratio and heat-set (details of processing varies with individual producers). Liu
(2016) employed wide angle x-ray scattering (WAXS), small angle x-ray scattering
(SAXS), and DSC for determining fiber morphology. They conclude that HMLS
has higher crystallinity and more perfect crystals than HMLE. Comparison of the
DSC traces of HMLS and HMLE, shown in Fig. 8.10, supports these conclusions,
although the detailed X-ray study performed provided more detailed morphological
Figure 8.10 Comparison of the melting behavior of HMLS and HMLE PET industrial yarns
(Liu et al., 2016). HMLE, High modulus low elongation; HMLS, high modulus low
shrinkage; PET, poly(ethylene terephthalate).
146
Thermal Analysis of Textiles and Fibers
information and supports the importance of the use of a variety of techniques when
studying the details of fiber structure.
Yoon (Yoon et al., 2017) showed how a modified PET drawing process
increased the crystallinity of PET industrial yarns and utilized DSC to measure
crystal-related changes. Potter (2012) published a review on the use of DSC and
TGA for the identification of textile fibers, including PET and other polyesters.
Gashti (Gashti and Navid, 2012) used DSC, TGA, and dynamic mechanical thermal analysis (DMTA) to study the effect of silicone emulsion softeners on the thermal and flammability properties of the resulting textile fibers. Their findings show
that the application of the silicone coating leads to an increased rate of thermal degradation and increased flammability of the resulting textile. Telli and Kale (2011)
examined the effect of ZnO nanoparticles, added to impart antibacterial properties
to the PET fiber, on the thermal properties of the resulting fiber. Fig. 8.11 shows
the changes in crystallization behavior as monitored by DSC, proving that the ZnO
is acting as a nucleating agent for the PET, as evidenced by the increase of the crystallization temperature as a function of ZnO content.
In a DSC and TGA study of polyester containing a Zn ion and phosphinic acid
compound flame retardant, Liu et al. concluded that the flame retardant enhanced
Figure 8.11 Changes in PET fiber as a function of ZnO concentration, MB signifies the PET
master batch used to spin fibers (Telli and Kale, 2011). PET, Poly(ethylene terephthalate).
Polyester fibers
147
the thermal stability of the resulting resin (Liu et al., 2014). Chen et al. (2014) studied the thermal conversion mechanism of polyester fiber (Dacron) and a variety of
other materials by TGA. A study of PET fiber cross-linking by trimethlylolpropane
and electron beam radiations was studied by DSC and TGA and the impact on
structure and properties was monitored (Zhu et al., 2017). In contrast to the fundamental studies of the materials science of PET fiber in the last century, work in the
21st century highlights the utility of thermal analysis in the study of specific PET
fiber products. Although the focus of the work has changed, the use of thermal
analysis techniques to elucidate the chemical and morphological impact on polyester fiber performance continues to be a basic tool in the study of presently ubiquitous polyester fiber products.
8.10
Other polyester fibers, polypropylene
terephthalate, poly(butylene terephthalate),
polyethylene naphthalate
Most of the work done on the thermal analysis of polyesters, other than PET, is
focused more on plastic than on fiber applications. In an article published in the
online New Fibres magazine in 2000, Brown et al. (2000) show TMA data for PTT
spunbond fibers. Thermal analysis of PEN fibers was treated in detail by Menczel
in 1996 (Jaffe et al., 1997). In general, any data relevant to molecularly oriented
structure of PPT, PBT, or PEN will relate to the expected thermal behavior of these
polymers in fiber form.
8.11
Conclusion
Polyester fiber thermal analysis has been extensively studied over the past 50 years,
resulting in a rich literature that utilizes thermal analysis techniques to develop
processstructureproperty relationships as well as to measure properties directly
related to polyester fiber and fiber-based product performance. The power of thermal analysis to provide a basic understanding of the mechanisms and processes,
which allow for the huge volume production of commercially important fiber
products, cannot be overstated.
References
Berndt, H.-J., Bossman, A., 1976. Polymer 17 (3), 241245.
Brown, H., Casey, P., Donahue, J., 2000. NF New Fibres Magazine. On-line.
148
Thermal Analysis of Textiles and Fibers
Burghardt, W., Vom, O.H., 1974. Process for the Manufacture of Fibers from High
Molecular Weight Linear Polyethylene Terephthalate, US 3803284 A.
Callender, D.G., 1985. Polym. Sci. Eng. 25, 453457.
Chang, Y., et al., 2004. Macromol. Mater. Eng. 289 (8), 703707.
Chen, et al., 2014. Fuel 137, 7784.
Cheremisinoff, N.P., 1989. Handbook of Polymer Science and Technology, Volume 1:
Synthesis and Properties. Marcel Dekker.
Davis, G.W., Hill, E.E., 1982. Kirk-Othmer Encyclopedia of Chemical Technology, Third
ed. Wiley, p. 535.
Easley, W.K., (1959). Canadian Patent 573,301 to Chemstrand Corp.
East, A.J., 2004. Kirk-Othmer Encyclopedia of Chemical Technology, vol. 5. Wiley.
Eichhorn, S., Hearle, W.S., Jaffe, M., Kikutani, T., 2009. Handbook of Textile Fiber
Structure, vol. 1. Woodhead Publishing, Oxford, Cambridge, New Delhi.
French Patent, 1959. French Patent 1,169,659 to I.C.I.
Gashti, M.P., Navid, M.Y., 2012. J. Appl. Polym. Sci. 125 (2), 14301438.
Heuvel, H.M., Husiman, R., 1978. J. Appl. Polym. Sci. 22 (8), 22292243.
Jaffe, M., Menczel, J.D., Bessey, W.E., 1981. Fibers. In: Turi, E. (Ed.), Thermal
Characterization of Polymer Materials. Academic Press, New York.
Jaffe, M., Menczel, J.D., Bessey, W.E., 1997. Fibers. In: Turi, E. (Ed.), Thermal
Characterization of Polymer Materials, second ed. Academic Press, San Diego, CA.
Kitano, et al., 1995. Polymer 36, 10.
Liu, Y., et al., 2014. Fangzhi Xuebao 35, 1216.
Liu, Y., et al., 2016. Polymer 105, 157166.
McIntyre, J.E., 2005. Synthetic Fibers: Nylon, Polyester, Acrylic, Polyolefin. Woodhead
Publishing.
Miller, R.W., Murayama, T., 1984. J/ Appl. Polym. Sci. 29 (3), 933939.
Nichols, C.S., Moore, T.C., Edwards, W.L., 1996. Method of Post-Polymerization
Stabilization of High Activity Catalysts in Continuous Polyethylene Terephthalate
Production, US Patent 5898058 A.
Paszun, D., Spychaj, T., 1997. Ind. Eng. Chem. Res. 46 (4), 13731383.
Potter, C., 2012. AATCC Rev. 12 (5), 3945.
SCI, 1999. Process Economics Program Report, 227, SCI International Reports.
Reese, G., 2003. Encyclopedia of Polymer Science and Technology, third ed. Wiley,
pp. 652678.
Scheirs, J., Long, T., 2005. Modern Polyesters: Chemistry and Technology of Polyesters and
Copolyesters. Wiley.
Sorenson, F., et al., 2001. Preparative Methods of Polymer Chemistry. Wiley.
Statista, 2019. Industrieverereingung Chemiefase, August, 2018.
Telli, M., Kale, R., 2011. Adv. Appl. Polym. Res. 2 (4), 491502.
Valk, G., Jellinek, G., Schroder, U., 1980. Tex. Res. J 50 (1), 4654.
Ward, I., et al., 2000. Solid Phase Processing of Polymers. Hansa Verlag.
Wilfong, R.E., 1961. J. Polym. Sci. 54, 388.
Yildirim, K., Ulcay, Y., 2014. e-Polymer 14 (2), 12.
Yoon, J., et al., 2017. Polym. Sci. Eng. 57 (2), 1290.
Zhu, S., Shi, M., Tian, M., Qu, L., Chen, G., 2018. Effects of irradiation on polyethyleneterephthalate (PET) fibers impregnated with sensitizer. J. Text. Inst. 109 (3), 294299.
Polyester fibers
149
Further reading
Lewin, M., 2006. Handbook of Fiber chemistry, third ed. CRC Press.
Menczel, J.D., Prime, B., 2009. Thermal Analysis of Polymers: Fundamentals and
Applications. Wiley.
Saw, C.K., Menczel, J.D., Choe, E. W., Hughes, O. R. (1997), SPE/ANTEC Proceedings,
610.
Towe, T., 2009. Interior Textiles Design and Developments. Woodhead Publishing.