Sunteți pe pagina 1din 16

Polyacrylonitrile

Polyacrylonitrile (PAN) is a general name for polymers consisting of at least 85% of


acrylonitrile monomer.

From: Advances in the Dyeing and Finishing of Technical Textiles, 2013

Related terms:

Nanofibers, Electrospinning, Nanoparticles, Carbon Fiber, Monomers, Polyamides,


Electrospun, Nanofiber

View all Topics

Learn more about Polyacrylonitrile

Polyacrylonitrile fibers
Bhupender S. Gupta, Mehdi Afshari, in Handbook of Properties of Textile and
Technical Fibres (Second Edition), 2018

15.5.3 Melt spinning


Polyacrylonitrile degrades before reaching its melting point (320°C). Water, on
the other hand, has a plasticizing effect on acrylic polymer and blocks the polar
interactions between nitrile groups, and causes a lowering of melting point ( 180°C
at 33%W water), thus making melt spinning possible (Frushour, 1981). There are
several patents published on melt spinning of polyacrylonitrile (Daumit et al., 1991;
Daumit and Ko, 1990; Porosofu, 1980; Yu et al., 2014, 2016). One of these shows
successful melt spinning of terpolymer of AN/MA/MAA (93:55:1.5) in the presence
of 12%–28% water, with 4.5–5.4 g/dtex tenacity (Gupta and Chand, 1979). In new
patents (Yu et al., 2016; Min et al., 1992), imidazole ion liquids have been used
as a plasticizer for melt spinning. Adding a comonomer to polyacrylonitrile can
effectively reduce the melting point as reported by Min et al. (1992). Melt spun acrylic
precursor has also been converted to carbon fiber. However, more voids and broken
filaments have been observed in the carbon fiber made from this process compared
to those made from the wet or dry spun fibers (Grover et al., 1988). In another study,
melt spinning of mixture of polyacrylonitrile and propylene carbonate at 190°C was
reported (Atureliya and Bashir, 1993).

> Read full chapter

A brief description of textile fibers


Yimin Qin, in Medical Textile Materials, 2016

Polyacrylonitrile
Polyacrylonitrile (PAN) is a synthetic polymer with the linear formula (C3H3N)n.
Although it is thermoplastic it does not melt under normal conditions, since it
degrades before melting at above 300 °C. Almost all PAN resins are copolymers
made from mixtures of monomers, with acrylonitrile as the main component. It
is a versatile polymer used to produce a large variety of products, including fibers
for textiles, ultra-filtration membranes, hollow fibers for reverse osmosis, etc. PAN
fibers are the precursor in the production of high-quality carbon fibers.

> Read full chapter

Surface Modification of Polymers


P. Fabbri, M. Messori, in Modification of Polymer Properties, 2017

5.5.5 PAN Membranes


Polyacrylonitrile (PAN) ultrafiltration membranes incorporating polyacrylonitrile-
-graft-polyethylene oxide (PAN-g-PEO) as amphiphilic comb copolymer additives
have been proposed to prevent irreversible adhesion of bacteria (Adout et al., 2010;
Asatekin et al., 2007).

In a very similar way, antifouling poly(vinylidene fluoride) ultrafiltration mem-


branes containing an amphiphilic comb polymer consisting of poly(vinylidene fluo-
ride-co-chlorotrifluoroethylene) [P(VDF-co-CTFE)] main chains and poly(oxyethylene
methacrylate) (POEM) side chains have been reported (Koh et al., 2010).

> Read full chapter


Developments in fibers for technical
nonwovens
Y. Yan, in Advances in Technical Nonwovens, 2016

2.3.5.4 Polyacrylonitrile
Polyacrylonitrile (PAN) is produced by the additional polymerization of acrylonitrile;
always, the second and the third monomers are used for the modification of dye-
ability and spinnability. They can then be spun into fibers by dry or wet spinning
methods, such as how Orlon with a distinctive dumbbell cross section is spun by dry
process produced by DuPont, and Acrilan had circular cross section and is spun by
the wet extrusion technique produced by Monsanto. PAN fiber can also get the crimp
structure like wool by using the bicomponent spinning process in fiber preparation.
Properties of PAN fibers are listed in Table 2.27.

Table 2.27. Properties of PAN fibers and their comparison with wool

Properties Fiber types


PAN staple fiber PAN filament Modified PAN staple Wool
fiber
Tenacity Dry 2.2–4.8 2.8–5.3 1.7–3.5 0.8–1.5
(cN/dtex)
Wet 1.7–3.9 2.6–5.3 1.7–3.5 0.7–1.4
Elongation (%) Dry 25–50 12–20 25–45 25–35
Wet 25–60 12–20 25–45 25–50
(Wet/dry strength) 80–100 90–100 90–100 76–96
(%)
Knot strength 1.4–3.1 2.6–7.1 1.3–2.5 0.7–1.2
(cN/dtex)
Hooking strength 1.6–3.4 1.7–3.5 1.4–2.5 0.7–1.3
(cN/dtex)
Initial Modulus 22–54 35–75 18–48 9.7–22
(cN/dtex)
Elastic recovery 90–95 70–95 85–95 98
rate (%) (elonga-
tion at 3%)
Moisture regain Commercial 2 2 15
(%)
standard state (20°C, 1.2–2.0 0.6–1.0 16
relative humidity, 65%)
Heat resistance (°C) – 150 100°C harden
Softening point (°C) 190–240 Not obvious 300°C carbonization
Melting point (°C) Not obvious – 130
Decomposition tem- 327 – –
perature (°C)
Glass transition tem- 80, 140 – –
perature (°C)
60 60 20
Light fastness (residual
strength for 12 months
exposure) (%)
Flammability (limiting 18.2 26.7 24–25
oxygen index) (%)
Acid resistance 35%a HCL, 65%a H2- 35%a HCL, 70%a H- Other hot acid resis-
SO4, 45%a HNO3 No 2SO4 No effect in tance, except H2SO4
effect in strength strength
Alkali resistance In 50%a NaOH and The same as PAN Bad alkali resistance,
28%a NH3·H2O, in- shrinking in dilute alkali
tensity of little decline
Solvent resistance Does not dissolve in Soluble in acetone, Does not dissolve in sol-
solvents vents
Bleaching resistance NaClO2 and H2O2 re- The same as PAN SO2 and H2O2 resis-
sistance tance
Abrasive resistance Good Good Common
Resistant to fungal Fungal resistance, not Fungal resistance, not Fungal resistance, dam-
damaged by worms damaged by worms aged by worms
Electrical insulating Dielectric constant: 6.5 Dielectric constant: 4.5 Specific resistance: 5 ×-
property (20°C, relative specific resistance: 2 ×-  108 Ω cm
humidity, 65%)  104 Ω cm
Dyeability Dispersed dyesDationic Dispersed dye cationic Acid dyesMordant dyes
dyesAcid dyes dye indigoid dyes

a All are mass fractions of relevant material.

Li G. Polymer material processing technology. Beijing: China Textile Press; 2010. p.


167.

> Read full chapter

Materials and technologies for


rechargeable lithium–sulfur batteries
N. Azimi, ... Z. Zhang, in Rechargeable Lithium Batteries, 2015

Sulfurized polyacrylonitrile
Polyacrylonitrile (PAN) is an excellent precursor for the conductive polymer of
sulfur–polymer composites. An in situ dehydrogenation, cyclization, and sulfuriza-
tion process has been employed to produce the so-called sulfurized polyacryloni-
trile (SPAN). The SPAN materials have comparable specific capacities with today’s
state-of-the-art S–C composite, and are compatible with the Li-ion battery elec-
trolytes based on LiPF6 salt and carbonate-based solvents [31,32]. The rate capability
of SPANs can be further improved by the incorporation of MWCNTs, in which
the MWCNTs enhance the structural stability and electronic conductivity of SPANs
[33,34]. In the same line of work, a pyrolyzed PAN–sulfur–graphene nanosheet
(pPAN-S-GNS) composite was prepared by impregnating sulfur into a PAN–GNS
composite synthesized by in situ polymerization of acrylonitrile and chemical reduc-
tion of graphene oxide (Figure 5.3) [35]. With 4 wt.% GNS added, the composite
showed a specific capacity of 800 mA h/g at relatively high C-rates (up to 6 °C) and
a 99.9% of coulombic efficiency. The excellent performance is attributed to the
three-dimensional GNS networks that enhance electronic conductivity and facilitate
distribution of the active material in the composite.

Figure 5.3. Schematic diagram of the in situ polymerization and synthesis of the
pPAN-S/GNS composite, in which the insets are cross-sectional views of the sam-
ples.

Reproduced from Ref. [35]. Copyright RSC Publishing. Reproduced with permission.

> Read full chapter

Textile Fibers: A Comparative Overview


J.W.S. Hearle, in Encyclopedia of Materials: Science and Technology, 2001

5.1 Polyacrylonitrile (Acrylic)


(PAN) was the third major synthetic fiber of the 1950s, but has not achieved the
wide usage of nylon and polyester. Apart from providing precursor yarns for carbon
fibers, PAN is almost entirely produced as staple fiber. The major uses are in warm,
bulky fabrics, such as knitwear, blankets, and pile fabrics, as an alternative to wool.
Its synthetic-fiber carpet market has been lost to nylon.

The basic chemical formula of PAN is


6.

In order to make dyeing possible, acrylic fibers (specified as >85% PAN) are copoly-
mers with added reactive groups. Acrylic fibers can be melt spun, but are usually
produced from solution by wet or dry spinning. The diversity of materials and
processes leads to differences in properties in fibers from different manufacturers
and in different options.

Acrylic fibers are weaker than nylon or polyester, yield at 2% extension, and have
poor recovery from higher extensions. Breaks are granular. They are regarded as
quasicrystalline, and have a major second-order transition near 100°C. Fibers which
have been highly stretched, or broken, show a high shrinkage when heated. Blending
of no-shrink or preshrunk and high-shrink fibers is a way of making bulky yarns.

> Read full chapter

The use of enzymatic techniques in the


finishing of technical textiles
R. Paul, E. Genescà, in Advances in the Dyeing and Finishing of Technical Textiles,
2013

8.3.3 Polyacrylonitrile
Polyacrylonitrile (PAN) is a general name for polymers consisting of at least 85% of
acrylonitrile monomer. Typically it consists of 89–90% acrylonitrile, 4–10% non-ion-
ic co-monomer (e.g. vinyl acetate) and around 1% ionic co-monomer containing
a sulpho (SO3H) or sulphonate (OSO3H) group (Guebitz and Cavaco-Paulo, 2008).
About 2.73 million tons of PAN per year is produced worldwide. There is an industrial
demand to improve moisture uptake, dyeability with ionic dyes and feasibility of
finishing processes that maintain its mechanical properties (Fischer-Colbrie et al.,
2007). Conventional chemical methods have not been very successful. High temper-
atures, aggressive chemicals and high concentrations of dimethyl sulphoxide have
led to unwanted changes in the fibre properties and yellowing of the fabrics, while
milder conditions have resulted in less hydrophobic fabrics. Selective enzymatic
hydrolysis of surface nitrile groups of PAN seems to be an interesting alternative
(Tauber et al., 2000).

Nitrilases catalyse the hydrolysis of a nitrile directly to the corresponding acid, while
the nitrile hydratase/amidase enzyme systems catalyse the hydrolysis in two steps,
first obtaining an amide and then the corresponding acid (Fischer-Colbrie et al.,
2007). Nitrile hydrolysing enzymes have been isolated from many organisms such
as Rhodococcus rhodochrous and Micrococcus luteus BST20. In the cases mentioned
above, amides have been detected during the hydrolysis, showing that the enzymatic
action is being developed by a nitrile hydratase/amidase system (Tauber et al., 2000;
Fischer-Colbrie et al., 2007). The hydrolysis of PAN fibres to the corresponding
amides with a nitrile hydratase from Brevibacterium imperiale and Corynebacterium
nitrilophilus was also studied, leading to improved dyeing properties (Battistel et al.,
2001).

In addition to nitrile hydrolysing enzymes, esterases can also be used for the
modification of PAN copolymers (Guebitz and Cavaco-Paulo, 2003). Cutinase has
been proved to be useful for the hydrolysis of the co-monomer vinyl acetate, yielding
acetic acid and leaving vinyl alcohol at the fibre surface (Silva et al., 2005c).

The enzyme reaction on PAN is restricted by several factors related to the properties
of the polymer, which leads to a limited hydrolysis. A maximum of 16% surface nitrile
hydrolysis has been reported (Tauber et al., 2000).

> Read full chapter

Exceptional CO2 capture via polymeric


materials
Hasmukh A. Patel, ... Mert Atilhan, in Proceedings of the 3rd Gas Processing
Symposium, 2012

3 Results
Polyamidoxime (PAO) has been prepared with the following procedure;

Polyacrylonitrile (1 mol) was dispersed in water followed by the addition of hydrox-


ylamine solution (50 wt. % in H2O) (1.1 mol) with uninterrupted agitation at 70 C
for 4h. The reaction mixture was centrifuged to isolate the off-white colored product
and dried under vacuum. Yield: 74 %. Infrared studies performed on the polymer
shows that PAO exhibits bands at 3450, 3353 cm−1 (N–H stretch), 3200 cm−1 (O–H
stretch), 1390 cm−1 (C–N stretch), 1654 cm−1 (C=N stretch), 931 cm−1 (N–O stretch).
Surface area was measured as 8.674 m2/g using BET analysis.

Gravimetric carbon dioxide adsorption capacities of various amidoximes were stud-


ied at two different temperatures (43° C and 70° C) up to 180 bar using the high-pres-
sure magnetic suspension balance (MSB) sorption device made by Rubotherm
Präzisionsmesstechnik GmbH.6 MSB apparatus was kept at 100 °C up to 350 bars
with two different operating positions. First, the measurement cell was filled with
carbon dioxide at the sorption measurement position. MSB recorded the change in
weight of the sample placed in the holder as the high pressure gas adsorbed on it.
The second measurement position was used to measure the in-situ density of the
high-pressure gas required to calculate the amount of the adsorbed gas on the solid
sample in the high-pressure cell. The data indicates that the adsorption isotherms
of amidoximes exhibit “Type IV” behavior type, in which multilayer adsorption are
observed after the completion of the first layer. Among the the amidoximes studied,
the adsorption isotherm of PAO possesses a clear distinction between the first layer
(0-60 bar region) and multilayer (60-180 bar region) adsorption indicating that the
adsorbate-adsorbent interactions present in this amidoxime is the strongest.

In order to assess chemical affinity of polymeric amidoxime to CO2, adsorption per


surface area is chosen to be a better fit since interaction of a guest gas molecule with
the high molecular weight solid sorbent is predominantly related to heterogeneous
interfaces and per weight arguments fall short of explaining the chemistry of sorp-
tivity in macromolecules (Figure 3). PAO gave adsorptions 0.53 and 0.41 mmol(CO2)
m−2 at 43 C and 70 C indicating significant chemical interaction between the gas
molecules and the sorbent, as opposed to the highly porous, inert Norit RB3.

Figure 3. CO2 adsorption isotherms (mmol m−2) for PAO () and Activated Charcoal
Norit RB3 () at 43 C.

Figure 2. CO2 adsorption isotherms (mmol/m2) for PAO at 43 C and 70 C.


> Read full chapter

FRACTURE OF CARBON FIBERS


J.G. Lavin, in Fiber Fracture, 2002

PAN-BASED CARBON FIBERS


Polyacrylonitrile (PAN) fibers are made by a variety of methods. The polymer is made
by free-radical polymerization either in solution or in a solvent-water suspension.
The polymer is then dried and re-dissolved in another solvent for spinning, either
by wet-spinning or dry-spinning. In the wet-spinning process the spin dope is
forced through a spinneret into a coagulating liquid and stretched, while in the
dry-spinning process the dope is spun into a hot gas chamber, and stretched. For
high-strength carbon fibers, it is important to avoid the formation of voids within
the fiber at this step. Dry-spun fibers are characterized by a ‘dog-bone’ cross-section,
formed because the perimeter of the fiber is quenched before much of the solvent
is removed. The preferred process for high-strength fiber today is wet-spinning.
Processes for melt-spinning PAN plasticized with water or polyethylene glycol have
been developed, but are not practiced commercially. A significant improvement in
carbon fiber strength was obtained by Moreton and Watt (1974) who spun the PAN
precursor under clean room conditions. The strength of fibers spun in this way and
subsequently heat treated was found to improve by >80% over conventionally spun
fibers. The mechanism is presumed to be removal of small impurities which can act
as crack initiators. This technology is believed to be critical for production of high
strength fibers such as Toray's T800 and T1000.

Initially, commercial PAN-based carbon fibers were made from the polymers devel-
oped for textile applications. However, these fibers were neither very stiff nor strong.
Development efforts over the 1960s and 1970s focused on increasing molecular
weight, introducing co-monomers to assist processing, and eliminating impurities
which limited mechanical strength. The chemistry of conversion of PAN to carbon
is quite complex, and the interested reader is referred to an excellent treatment in
Peebles (1994). The critical steps are outlined below.

The first critical step in making carbon fiber from PAN fiber is causing the pendant
nitrile groups to cyclize, as illustrated in Fig. 8. This process is thermally activated
and is highly exothermic. The activation temperature is influenced by the type and
amount of co-monomer used. It is also important to keep the fiber under tension
in this process, and indeed, during the whole conversion process. The next step is
to make the fiber infusible: this is accomplished by adding oxygen atoms to the
polymer, again by heating in air. The reaction is diffusion limited, requiring exposure
times of tens of minutes. When about 8% oxygen by weight has been added, the
fiber can be heated above 600°C without melting. When the fiber is heated above
this temperature, the processes of decyanization and dehydrogenation take place,
and above 1000°C large aromatic sheets start to form, as illustrated in Fig. 9.

Fig. 8. PAN-based carbon fiber chemistry: cyclization and oxidation.


Fig. 9. PAN-based carbon fiber chemistry: carbonization.

The weight loss experienced in the production of carbon fibers from PAN precursor
is approximately 50%. Guigon et al. (1984) showed that this leads to a structure
containing many longitudinal voids, as shown in Fig. 10, and a density of 1.8 g/cm3,
compared with 2.28 g/cm3 for pure graphite, and 2.1 for pitch-based carbon fibers.
Boyes and Lavin (1998) showed evidence for the polymeric nature of the fiber in the
fracture surface shown in Fig. 11. The fibrils are evident on the wall of the fiber. An
enlargement of the fracture surface in Fig. 12 shows fibrils at the nanometer scale.
The results of a remarkable experiment by Kwizera et al. (1982) are shown in Fig. 13. A
Celion GY-70 fiber was fractured in vacuum, and exploded into microfibrils roughly
100 nm in diameter, further confirming the fibrillar nature of the PAN-based fiber.
Fig. 10. Model of microtexture of PAN-based carbon fiber.

(Copyright 1984, reproduced with permission from Elsevier Science.)

Copyright © 1984
Fig. 11. PAN-based carbon fiber fracture surface.

Fig. 12. Enlargement of fibrils in PAN-based carbon fiber fracture surface.


Fig. 13. PAN-based carbon fiber fractured in vacuum. The fibrils are approximately
100 nm in diameter.

(Copyright 1982, reproduced with permission from Elsevier Science.)

Copyright © 1982

> Read full chapter

Specialty Polymers & Polymer Process-


ing
Joseph Zimmerman, in Comprehensive Polymer Science and Supplements, 1989

8.2.4 Polyacrylonitrile and Polyalkenes


Polyacrylonitrile (PAN), linear polyethylene (PE) and isotactic polypropylene (PP) are
also significant polymers for the fiber industry. Their monomers are all H2CCHX,
where X is CN, H and Me respectively for the three polymers. The principles
of their polymerization are covered elsewhere in these volumes. However some
key points will briefly be mentioned. They all polymerize via addition polymeriza-
tion. The polymerization of PAN is a free radical addition polymerization initiated
by such materials as azobis(isobutyronitrile), benzoyl peroxide and persulfates.83
Comonomers are frequently used to enhance solubility in spinning solvents such as
dimethylformamide and the dyeability of fibers. Acrylic fibers contain at least 85% of
acrylonitrile (AN), while modacrylic fibers contain less than 85% but more than 35%
of AN. Some expressions which have been proposed for the relationship between
[ ] and molecular weight for PAN are83 [ ]= 2.33 × 10−4 and [ ] = 2.43 × 10−4 so that
M̄w = 100 000 corresponds to [ ] = 1.31. Typical fibers have [ ] values of about 1.5.

Polyethylene and polypropylene are polymerized anionically with the help of Zeigler
type catalysts,84,85 which are generally based on TiCl3 and metal alkyls or alkyl halides.
The linear polymers produced by this type of process tend to be relatively highly
crystalline with densities in the range of 0.95 for PE and 0.90 for PP. The molecular
weight distributions of typical polyalkenes tend to be much broader than those
encountered with condensation polymers with M̄w/M̄n values ranging from three to
15. In recent work with PE, polymers with M̄w as high as 4 × 106 have been spun
into fibers. Based on various data in the literature,85–87 the relationship [ ] = 6.3
× 10−4 seems to fit much of the data so that [ ] = 10 corresponds to M̄w = 106.
Some differences between the various sources may be a result of differences in the
molecular weight distributions. A relationship proposed for polypropylene84 for inh
(0.1% in decalin at 135 °C) is inh = 1.00 × 10−4 so that inh = 6 corresponds to
M̄w = 106. This relationship seems to apply to data given in ref. 86 as well. Another
relationship85 gives [ ] = 1.62 × 10−4. M̄w/M̄n for commercial polypropylene is about
1.3.85

The melt viscosity of high density polyethylene at a given molecular weight is lower
than that for polyamides because of the absence of hydrogen bonding. However,
higher molecular weights are generally used for PE, in part because molecular
weight distributions are broader. Non-Newtonian behavior begins at shear stresses
of about 103 Pa or less.89 The activation energy of viscosity90,91 is about 28–29 kJ mol−1
(6.8–7 kcal mol−1), lower than that for the polyamides, again because of the absence
of hydrogen bonding. While a power law relationship is generally used for shear
stress vs. shear rate data in the non-Newtonian region, an alternative which has been
used90 over the entire range of shear rates is to plot log( / 0) vs. shear stress to obtain
a straight line. Polypropylene also becomes non-Newtonian at about 103 Pa shear
stress but the onset stress decreases and the degree of viscosity reduction increases
as the molecular weight distribution broadens.89,92 As expected, viscosity increases as
about the 3.5 power of molecular weight (M̄w or M̄v)92 as it does for polyethylene and
other linear polymers. The activation energy for viscosity of PP is about 46 kJ mol−1
(11 kcal mol−1).91 As an example,89 the absolute value of zero shear melt viscosity at
200 °C for high density polyethylene with M̄w = 1.68 × 105(M̄w/M̄n = 84) is 1.9 × 105 Pa
s, while that for polypropylene with M̄w = 4.44 × 105(M̄w/M̄n = 10.4) is 5.7 × 103 Pa s.

> Read full chapter


ScienceDirect is Elsevier’s leading information solution for researchers.
Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

S-ar putea să vă placă și