Sei sulla pagina 1di 73

List Learning Objective

 To discuss issues relating to the disposal and recycling of polymers and the
conservation of resources used in polymer manufacture

 To relate the properties of polymers to their molecular structure, and make


predictions about the properties of a polymer

 To explain the properties of condensation polymers in terms of intermolecular


attractions eg H bonding in nylon or permanent dipole - permanent dipole
interactions in PET

 To show awareness of the ways that chemists can modify the properties of a
polymer by physical and chemical means.

 To distinguish between addition and condensation polymerization.

List Learning Outcomes

 Understand what polymers are made up of, their properties, and basic polymer
nomenclature
 Identify the different polymerisation and polymer analytical techniques
 Differentiate between commodity and engineering thermoplastics
 Appreciate how polymers are processed for specific uses
 Suitable polymeric material for a daily use product
Keywords in Thermochesmistry

 Polymerisation

 DNA Polymerase

 Propagation

 Terminal pathway

 Linear Polymer

 Ester linkages

 Quasi-thermoplastic

 Cross link process

 Polyisoprne

 Delocalisation

Subtopics

 Synthetic Polymers

 Addition Polymerization Mechanisms

 Coordination Polymerization

 Condensation Polymerization

 Thermoplastic and Thermosetting Materials

 Natural Rubber
Introduction

Prior to the early1920's, chemists doubted the existence of molecules having

molecular weights greater than a few thousand. This limiting view was

challenged by Hermann Staudinger, a German chemist with experience in

studying natural compounds such as rubber and cellulose. In contrast to the

prevailing rationalization of these substances as aggregates of small molecules,

Staudinger proposed they were made up of macromolecules composed of

10,000 or more atoms. He formulated a polymeric structure for rubber, based on

a repeating isoprene unit (referred to as a monomer). For his contributions to

chemistry, Staudinger received the 1953 Nobel Prize. The terms polymer and

monomer were derived from the Greek roots poly (many), mono (one) and meros

(part).

Recognition that polymeric macromolecules make up many important natural

materials was followed by the creation of synthetic analogs having a variety of

properties. Indeed, applications of these materials as fibers, flexible films,

adhesives, resistant paints and tough but light solids have transformed modern

society. Some important examples of these substances are discussed in the

following sections.
Historical development

Starting in 1811, Henri Braconnot did pioneering work in derivative cellulose


compounds, perhaps the earliest important work in polymer science. The development of
vulcanization later in the nineteenth century improved the durability of the natural
polymer rubber, signifying the first popularized semi-synthetic polymer. In 1907, Leo
Baekeland created the first completely synthetic polymer, Bakelite, by reacting phenol
and formaldehyde at precisely controlled temperature and pressure. Bakelite was then
publicly introduced in 1909.

Despite significant advances in synthesis and characterization of polymers, a correct


understanding of polymer molecular structure did not emerge until the 1920s. Before
then, scientists believed that polymers were clusters of small molecules (called colloids),
without definite molecular weights, held together by an unknown force, a concept known
as association theory. In 1922, Hermann Staudinger proposed that polymers consisted of
long chains of atoms held together by covalent bonds, an idea which did not gain wide
acceptance for over a decade and for which Staudinger was ultimately awarded the Nobel
Prize. Work by Wallace Carothers in the 1920s also demonstrated that polymers could be
synthesized rationally from their constituent monomers. An important contribution to
synthetic polymer science was made by the Italian chemist Giulio Natta and the German
chemist Karl Ziegler, who won the Nobel Prize in Chemistry in 1963 for the development
of the Ziegler-Natta catalyst. Further recognition of the importance of polymers came
with the award of the Nobel Prize in Chemistry in 1974 to Paul Flory, whose extensive
work on polymers included the kinetics of step-growth polymerization and of addition
polymerization, chain transfer, excluded volume, the Flory-Huggins solution theory, and
the Flory convention.

Synthetic polymer materials such as nylon, polyethylene, Teflon, and silicone have
formed the basis for a burgeoning polymer industry. These years have also shown
significant developments in rational polymer synthesis. Most commercially important
polymers today are entirely synthetic and produced in high volume on appropriately
scaled organic synthetic techniques. Synthetic polymers today find application in nearly
every industry and area of life. Polymers are widely used as adhesives and lubricants, as
well as structural components for products ranging from children's toys to aircraft. They
have been employed in a variety of biomedical applications ranging from implantable
devices to controlled drug delivery. Polymers such as poly(methyl methacrylate) find
application as photoresist materials used in semiconductor manufacturing and low-k
dielectrics for use in high-performance microprocessors. Recently, polymers have also
been employed as flexible substrates in the development of organic light-emitting diodes
for electronic displays.
Synthetic polymers are often referred to as "plastics", such as the well-known
polyethylene and nylon. However, most of them can be classified in at least three main
categories: thermoplastics, thermosets and elastomers. Formed from hydrocarbons,
hydrocarbon derivatives, or sometimes from silicon, polymers are the basis not only for
numerous natural materials, but also for most of the synthetic plastics that one encounters
every day. Polymers consist of extremely large, chain-like molecules that are, in turn,
made up of numerous smaller, repeating units called monomers. Chains of polymers can
be compared to paper clips linked together in long strands, and sometimes cross-linked to
form even more durable chains. Polymers can be composed of more than one type of
monomer, and they can be altered in other ways. Likewise they are created by two
different chemical processes, and thus are divided into addition and condensation
polymers. Among the natural polymers are wool, hair, silk, rubber, and sand, while the
many synthetic polymers include nylon, synthetic rubber, Teflon, Formica, Dacron, and
so forth. It is very difficult to spend a day without encountering a natural polymer—even
if hair is removed from the list—but in the twenty-first century, it is probably even harder
to avoid synthetic polymers, which have collectively revolutionized human existence.

They are not limited to having carbon backbones, elements such as silicon form familiar
materials such as silicones. Coordination polymers may contain a range of metals in the
backbone, with non-covalent bonding present.

Man-made polymers are used in a wide array of applications: food packaging, films,
fibers, tubing, pipes, etc. The personal care industry also uses polymers to aid in texture
of products, binding, and moisture retention (e.g. in hair gel and conditioners).

Laboratory synthesis

Laboratory synthetic methods are generally divided into two categories, step-growth
polymerization and chain-growth polymerization[4]. The essential difference between the
two is that in chain growth polymerization, monomers are added to the chain one at a
time only[5], whereas in step-growth polymerization chains of monomers may combine
with one another directly[6]. However, some newer methods such as plasma
polymerization do not fit neatly into either category. Synthetic polymerization reactions
may be carried out with or without a catalyst. Efforts towards rational synthesis of
biopolymers via laboratory synthetic methods, especially artificial synthesis of proteins,
is an area of intense research.
Biological synthesis

There are three main classes of biopolymers: polysaccharides, polypeptides, and


polynucleotides. In living cells, they may be synthesized by enzyme-mediated processes,
such as the formation of DNA catalyzed by DNA polymerase. The synthesis of proteins
involves multiple enzyme-mediated processes to transcribe genetic information from the
DNA to RNA and subsequently translate that information to synthesize the specified
protein from amino acids. The protein may be modified further following translation in
order to provide appropriate structure and functioning.

Modification of natural polymers

Many commercially important polymers are synthesized by chemical modification of


naturally occurring polymers. Prominent examples include the reaction of nitric acid and
cellulose to form nitrocellulose and the formation of vulcanized rubber by heating natural
rubber in the presence of sulphur.

Organic examples

As explained in the essay on Organic Chemistry, chemists once defined the term
"organic" as relating only to living organisms; the materials that make them up; materials
derived from them; and substances that come from formerly living organisms. This
definition, which more or less represents the everyday meaning of "organic," includes a
huge array of life forms and materials: humans, all other animals, insects, plants,
microorganisms, and viruses; all substances that make up their structures (for example,
blood, DNA, and proteins); all products that come from them (a list diverse enough to
encompass everything from urine to honey); and all materials derived from the bodies of
organisms that were once alive (paper, for instance, or fossil fuels).

As broad as this definition is, it is not broad enough to represent all the substances
addressed by organic chemistry—the study of carbon, its compounds, and their
properties. All living or once-living things do contain carbon; however, organic
chemistry is also concerned with carbon-containing materials—for instance, the synthetic
plastics we will discuss in this essay—that have never been part of a living organism.

It should be noted that while organic chemistry involves only materials that contain
carbon, carbon itself is found in other compounds not considered organic: oxides such as
carbon dioxide and monoxide, as well as carbonates, most notably calcium carbonate or
limestone. In other words, as broad as the meaning of "organic" is, it still does not
encompass all substances containing carbon.
Some of the most important polymers, according to monomer composition, are:

• Polyacrylates
• Polyamides
• Polyesters
• Polycarbonates
• Polyimides
• Polystyrenes

However, a polymer need not be wholly made from one class of monomer, in which case
it is classified as a copolymer.

A non-exhaustive list of ubiquitous materials includes:

• acrylonitrile butadiene styrene (ABS)


• polyacrylonitrile (PAN) or Acrylic
• polybutadiene
• poly(butylene terephthalate) (PBT)
• poly(ether sulfone) (PES, PES/PEES)
• poly(ether ether ketone)s (PEEK, PES/PEEK)
• polyethylene (PE)
• poly(ethylene glycol) (PEG)
• poly(ethylene terephthalate) (PET)
• polypropylene (PP)
• polytetrafluoroethylene (PTFE)
• styrene-acrylonitrile resin (SAN)
• poly(trimethylene terephthalate) (PTT)
• polyurethane (PU)
• polyvinyl butyral (PVB)
• polyvinylchloride (PVC)
• polyvinylidenedifluoride (PVDF)
• poly(vinyl pyrrolidone) (PVP)

Inorganic examples

• Polysiloxanes
• Polyphosphazenes

Brand names
These polymers are often better known through their brand names, for instance:

• Bakelite, i.e. phenol-formaldehyde resin


• Kevlar, Twaron, i.e. para-aramid
• Kynar, i.e. PVDF
• Mylar, i.e. polyethylene terephthalate film
• Neoprene i.e. Polychloroprene
• Nylon, i.e. polyamide 6,6
• Orlon, i.e. polyacrylonitrile
• Rilsan, i.e. polyamide 11 & 12
• Technora, i.e. copolyamid
• Teflon, i.e. PTFE
• Ultem, i.e. polyimide
• Vectran, i.e. aromatic polyamide
• Viton, i.e. poly-tetrafluoroethylene
• Zylon, i.e. poly-p-phenylene-2,6-benzobisoxazole (PBO)

Synthetic Biodegradable Polymers as Medical


Devices
In the first half of this century, research into materials synthesized from glycolic acid and
other -hydroxy acids was abandoned for further development because the resulting
polymers were too unstable for long-term industrial uses. However, this very instability
—leading to biodegradation—has proven to be immensely important in medical
applications over the last three decades. Polymers prepared from glycolic acid and lactic
acid have found a multitude of uses in the medical industry, beginning with the
biodegradable sutures first approved in the 1960s. Since that time, diverse products based
on lactic and glycolic acid—and on other materials, including poly(dioxanone),
poly(trimethylene carbonate) copolymers, and poly ( -caprolactone) homopolymers and
copolymers—have been accepted for use as medical devices. In addition to these
approved devices, a great deal of research continues on
polyanhydrides, polyorthoesters, polyphosphazenes, and
other biodegradable polymers.

A biodegradable intravascular stent prototype is molded


from a blend of polylactide and trimethylene carbonate.
Photo: Cordis Corp. Prototype Molded by Tesco
Associates, Inc.

Why would a medical practitioner want a material to degrade? There may be a variety of
reasons, but the most basic begins with the physician's simple desire to have a device that
can be used as an implant and will not require a second surgical intervention for removal.
Besides eliminating the need for a second surgery, the biodegradation may offer other
advantages. For example, a fractured bone that has been fixated with a rigid,
nonbiodegradable stainless implant has a tendency for refracture upon removal of the
implant. Because the stress is borne by the rigid stainless steel, the bone has not been able
to carry sufficient load during the healing process. However, an implant prepared from
biodegradable polymer can be engineered to degrade at a rate that will slowly transfer
load to the healing bone. Another exciting use for which biodegradable polymers offer
tremendous potential is as the basis for drug delivery, either as a drug delivery system
alone or in conjunction to functioning as a medical device.

Polymer scientists, working closely with those in the device and medical fields, have
made tremendous advances over the last 30 years. This article will focus on a number of
these developments. We will also review the chemistry of the polymers, including
synthesis and degradation, describe how properties can be controlled by proper synthetic
controls such as copolymer composition, highlight special requirements for processing
and handling, and discuss some of the commercial devices based on these materials.

POLYMER CHEMISTRY

Biodegradable
Polymer Melting Glass- Modulus Degradation polymers can be either
Point (°C) Transition (Gpa)a Time natural or synthetic. In
Temp (°C) (months)b general, synthetic
polymers offer greater
PGA 225—230 35—40 7.0 6 to 12 advantages than natural
LPLA 173—178 60—65 2.7 >24 materials in that they
can be tailored to give a
DLPLA Amorphous 55—60 1.9 12 to 16 wider range of
properties and more
PCL 58—63 (—65)— 0.4 >24 predictable lot-to-lot
(—60) uniformity than can
materials from natural
PDO N/A (—10)— 0 1.5 6 to 12
sources. Synthetic
PGA- N/A N/A 2.4 6 to 12 polymers also represent
TMC a more reliable source
of raw materials, one
85/15 Amorphous 50—55 2.0 5 to 6 free from concerns of
DLPLG immunogenicity.

75/25 Amorphous 50—55 2.0 4 to 5


DLPLG
65/35 Amorphous 45—50 2.0 3 to 4
Table I. Properties of
DLPLG
common biodegradable
50/50 Amorphous 45—50 2.0 1 to 2 polymers.
DLPLG
a Tensile or flexural modulus.
b Time to complete mass loss. Rate also depends on part
geometry.
The general criteria for selecting a polymer for use as a biomaterial is to match the
mechanical properties and the time of degradation to the needs of the application (see
Table I). The ideal polymer for a particular application would be configured so that it:

• Has mechanical properties that match the application, remaining sufficiently


strong until the surrounding tissue has healed.
• Does not invoke an inflammatory or toxic response.
• Is metabolized in the body after fulfilling its purpose, leaving no trace.
• Is easily processable into the final product form.
• Demonstrates acceptable shelf life.
• Is easily sterilized.

The factors affecting the mechanical performance of biodegradable polymers are those
that are well known to the polymer scientist, and include monomer selection, initiator
selection, process conditions, and the presence of additives. These factors in turn
influence the polymer's hydrophilicity, crystallinity, melt and glass-transition
temperatures, molecular weight, molecular-weight distribution, end groups, sequence
distribution (random versus blocky), and presence of residual monomer or additives. In
addition, the polymer scientist working with biodegradable materials must evaluate each
of these variables for its effect on biodegradation.1

Biodegradation has been accomplished by synthesizing polymers that have hydrolytically


unstable linkages in the backbone. The most common chemical functional groups with
this characteristic are esters, anhydrides, orthoesters, and amides. We will discuss the
importance of the properties affecting biodegradation later in the article.

The following section presents an overview of the synthetic biodegradable polymers that
are currently being used or investigated for use in wound closure (sutures, staples);
orthopedic fixation devices (pins, rods, screws, tacks, ligaments); dental applications
(guided tissue regeneration); cardiovascular applications (stents, grafts); and intestinal
applications (anastomosis rings). Most of the commercially available biodegradable
devices are polyesters composed of homopolymers or copolymers of glycolide and
lactide. There are also devices made from copolymers of trimethylene carbonate and -
caprolactone, and a suture product made from polydioxanone.

Polyglycolide (PGA). Polyglycolide is the simplest linear aliphatic polyester. PGA was
used to develop the first totally synthetic absorbable suture, marketed as Dexon in the
1960s by Davis and Geck, Inc. (Danbury, CT). Glycolide monomer is synthesized from
the dimerization of glycolic acid. Ring-opening polymerization yields high-molecular-
weight materials, with approximately 1—3% residual monomer present (see Figure 1).
PGA is highly crystalline (45—55%), with a high melting point (220—225°C) and a
glass-transition temperature of 35—40°C. Because of its high degree of crystallization, it
is not soluble in most organic solvents; the exceptions are highly fluorinated organics
such as hexafluoroisopropanol. Fibers from PGA exhibit high strength and modulus and
are too stiff to be used as sutures except in the form of braided material. Sutures of PGA
lose about 50% of their strength after 2 weeks and 100% at 4 weeks, and are completely
absorbed in 4—6 months. Glycolide has been copolymerized with other monomers to
reduce the stiffness of the resulting fibers.

Figure 1. Synthesis of polyglycolide (PGA).

Polylactide (PLA). Lactide is the cyclic dimer of lactic acid that exists as two optical
isomers, d and l. l-lactide is the naturally occurring isomer, and dl-lactide is the synthetic
blend of d-lactide and l-lactide. The homopolymer of l-lactide (LPLA) is a
semicrystalline polymer. These types of materials exhibit high tensile strength and low
elongation, and consequently have a high modulus that makes them more suitable for
load-bearing applications such as in orthopedic fixation and sutures. Poly(dl-lactide)
(DLPLA) is an amorphous polymer exhibiting a random distribution of both isomeric
forms of lactic acid, and accordingly is unable to arrange into an organized crystalline
structure. This material has lower tensile strength, higher elongation, and a much more
rapid degradation time, making it more attractive as a drug delivery system. Poly(l-
lactide) is about 37% crystalline, with a melting point of 175—178°C and a glass-
transition temperature of 60—65°C. The degradation time of LPLA is much slower than
that of DLPLA, requiring more than 2 years to be completely absorbed. Copolymers of l-
lactide and dl-lactide have been prepared to disrupt the crystallinity of l-lactide and
accelerate the degradation process.
Poly( -caprolactone). The ring-opening polymerization of -caprolactone yields a
semicrystalline polymer with a melting point of 59—64°C and a glass-transition
temperature of —60°C (see Figure 2). The polymer has been regarded as tissue
compatible and used as a biodegradable suture in Europe. Because the homopolymer has
a degradation time on the order of 2 years, copolymers have been synthesized to
accelerate the rate of bioabsorption. For example, copolymers of -caprolactone with dl-
lactide have yielded materials with more-rapid degradation rates. A block copolymer of -
caprolactone with glycolide, offering reduced stiffness compared with pure PGA, is being
sold as a monofilament suture by Ethicon, Inc. (Somerville, NJ), under the trade name
Monacryl.

Figure 2. Synthesis of poly( -caprolactone).

Poly(dioxanone) (a polyether-ester). The ring-opening polymerization of p-dioxanone


(see Figure 3) resulted in the first clinically tested monofilament synthetic suture, known
as PDS (marketed by Ethicon). This material has approximately 55% crystallinity, with a
glass-transition temperature of —10 to 0°C. The polymer should be processed at the
lowest possible temperature to prevent depolymerization back to monomer.
Poly(dioxanone) has demonstrated no acute or toxic effects on implantation. The
monofilament loses 50% of its initial breaking strength after 3 weeks and is absorbed
within 6 months, providing an advantage over Dexon or other products for slow-healing
wounds.
Figure 3. Synthesis of poly(dioxanone). -caprolactone).

Poly(lactide-co-glycolide). Using the polyglycolide and poly(l-lactide) properties as a


starting point, it is possible to copolymerize the two monomers to extend the range of
homopolymer properties (see Figure 4). Copolymers of glycolide with both l-lactide and
dl-lactide have been developed for both device and drug delivery applications. It is
important to note that there is not a linear relationship between the copolymer
composition and the mechanical and degradation properties of the materials. For
example, a copolymer of 50% glycolide and 50% dl-lactide degrades faster than either
homopolymer (see Figure 5). Copolymers of l-lactide with 25—70% glycolide are
amorphous due to the disruption of the regularity of the polymer chain by the other
monomer. A copolymer of 90% glycolide and 10% l-lactide was developed by Ethicon as
an absorbable suture material under the trade name Vicryl. It absorbs within 3—4 months
but has a slightly longer strength-retention time.

Figure 4. Synthesis of poly(lactide-co-glycolide). -caprolactone).

Figure 5. Half-life of PLA and PGA


homopolymers and copolymers implanted in rat
tissue. (Figure reproduced courtesy of Journal
of Biomedical Materials Research, 11:711,
1977.)

Copolymers of glycolide with trimethylene carbonate (TMC), called polyglyconate (see


Figure 6), have been prepared as both sutures (Maxon, by Davis and Geck) and as tacks
and screws (Acufex Microsurgical, Inc., Mansfield, MA). Typically, these are prepared
as A-B-A block copolymers in a 2:1 glycolide:TMC ratio, with a glycolide-TMC center
block (B) and pure glycolide end blocks (A). These materials have better flexibility than
pure PGA and are absorbed in approximately 7 months. Glycolide has also been
polymerized with TMC and p-dioxanone (Biosyn, by United States Surgical Corp.,
Norwalk, CT) to form a terpolymer suture that absorbs within 3—4 months and offers
reduced stiffness compared with pure PGA fibers.

Figure 6. Synthesis of polyglyconate.

Other Polymers under Development. Currently, only devices made from


homopolymers or copolymers of glycolide, lactide, caprolactone, p-dioxanone, and
trimethylene carbonate have been cleared for marketing by FDA. A number of other
polymers, however, are being investigated for use as materials for biodegradable devices.

In addition to their suitability for medical uses, biodegradable polymers make excellent
candidates for packaging and other consumer applications. A number of companies are
evaluating ways to make low-cost biodegradable polymers. One method is to bioengineer
the synthesis of the polymers, using microorganisms to produce energy-storing
polyesters. Two examples of these materials—polyhydroxybutyrate (PHB) and
polyhydroxyvalerate (PHV)—are commercially available as copolymers under the trade
name Biopol (Monsanto Co., St. Louis) and have been studied for use in medical devices
(see Figure 7). The PHB homopolymer is crystalline and brittle, whereas the copolymers
of PHB with PHV are less crystalline, more flexible, and easier to process. These
polymers typically require the presence of enzymes for biodegradation but can degrade in
a range of environments and are under consideration for several biomedical applications.

Figure 7. Molecular structure of two bioengineered polyesters that require specific


enzymes for biodegradation.
The use of synthetic poly(amino acids) as polymers for biomedical devices would seem a
logical choice, given their wide occurrence in nature. In practice, however, pure insoluble
poly(amino acids) have found little utility because of their high crystallinity, which
makes them difficult to process and results in relatively slow degradation. The
antigenicity of polymers with more than three amino acids in the chain also makes them
inappropriate for use in vivo. To circumvent these problems, modified "pseudo"
poly(amino acids) have been synthesized by using a tyrosine derivative. Tyrosine-derived
polycarbonates, for example, are high-strength materials that may be useful as orthopedic
implants. It is also possible to copolymerize poly(amino acids) to modify their properties.
The group that has been researched most extensively is the polyesteramides.

A Note on Nomenclature

A polymer is generally named based on the monomer it is synthesized from. For


example, ethylene is used to produce poly(ethylene). For both glycolic acid and lactic
acid, an intermediate cyclic dimer is prepared and purified, prior to polymerization.
These dimers are called glycolide and lactide, respectively. Although most references in
the literature refer to polyglycolide or poly(lactide), you will also find references to
poly(glycolic acid) and poly(lactic acid). Poly(lactide) exists in two stereo forms,
signified by d or l for dexorotary or levorotary, or by dl for the racemic mix.

The search for new candidate polymers for drug delivery may offer potential for medical
device applications as well. In drug delivery, the formulation scientist is concerned not
only with shelf-life stability of the drug but also with stability after implantation, when
the drug may reside in the implant for 1—6 months or more. For drugs that are
hydrolytically unstable, a polymer that absorbs water may be contraindicated, and
researchers have begun evaluating more hydrophobic polymers that degrade by surface
erosion rather than by bulk hydrolytic degradation. Two classes of these polymers are the
polyanhydrides and the polyorthoesters.

Figure 8. Molecular structure of poly(SA-HDA anhydride).

Polyanhydrides have been synthesized via the dehydration of diacid molecules by melt
polycondensation (see Figure 8). Degradation times can be adjusted from days to years
according to the degree of hydrophobicity of the monomer selected. The materials
degrade primarily by surface erosion and possess excellent in vivo compatibility. So far,
they have only been approved for sale as a drug delivery system. The Gliadel product,
designed for delivery of the chemotherapeutic agent BCNU in the brain, received
regulatory clearance from FDA in 1996 and is being produced by Guilford
Pharmaceuticals, Inc. (Baltimore).

Figure 9. Molecular structure of


poly(orthoester).

Polyorthoesters were first investigated in the


1970s by Alza Corp. (Palo Alto, CA) and SRI
International (Menlo Park, CA) in a search for new synthetic biodegradable polymers for
drug delivery applications (see Figure 9). These materials have gone through several
generations of improvements in synthesis, and can now be polymerized at room
temperature without forming condensation by-products. Polyorthoesters are hydrophobic,
with hydrolytic linkages that are acid-sensitive but stable to base. They degrade by
surface erosion, and degradation rates can be controlled by incorporation of acidic or
basic excipients.

PACKAGING AND STERILIZATION

Because biodegradable polymers are hydrolytically unstable, the presence of moisture


can degrade them in storage, during processing, and after device fabrication. In theory,
the solution for hydrolysis instability is simple: eliminate the moisture and thus eliminate
the degradation. However, because the materials are naturally hygroscopic, eliminating
water and then keeping the polymer free of water are difficult to accomplish. The as-
synthesized polymers have relatively low water contents, since any residual water in the
monomer is used up in the polymerization reaction. The polymers are quickly packaged
after manufacture—generally double-bagged under an inert atmosphere or vacuum. The
bag material may be polymeric or foil, but it must be highly resistant to water
permeability. To minimize the effects of any moisture present, the polymers are typically
stored in a freezer. Packaged polymers should always be at room temperature when
opened to minimize condensation, and should be handled as little as possible at ambient
atmospheric conditions. As expected, there is a relationship among biodegradation rate,
shelf stability, and polymer properties. For instance, the more hydrophilic glycolide
polymers are much more sensitive to hydrolytic degradation than are polymers prepared
from the more hydrophobic lactide.

Final packaging consists of placing the suture or device in an airtight, moistureproof


container. A desiccant can be added to further reduce the effects of moisture. Sutures, for
example, are wrapped around a specially dried paper holder that acts as a desiccant. In
some cases, the finished device may be stored at subambient temperature as an added
precaution against degradation.

Devices incorporating biodegradable polymers cannot be subjected to autoclaving, and


must be sterilized by gamma or E-beam irradiation or by exposure to ethylene oxide
(EtO) gas. There are certain disadvantages, however, to both irradiation and EtO
sterilization. Irradiation, particularly at doses above 2 Mrd, can induce significant
degradation of the polymer chain, resulting in reduced molecular weight as well as
influencing final mechanical properties and degradation times. Polyglycolide,
poly(lactide), and poly(dioxanone) are especially sensitive to ionizing radiation, and these
materials are usually sterilized by EtO for device applications. Because the highly toxic
EtO can present a safety hazard, great care must be taken to ensure that all the gas is
removed from the device before final packaging. The temperature and humidity
conditions should also be considered when submitting devices for sterilization.
Temperatures must be kept below the glass-transition temperature of the polymer to
prevent the part geometry from changing during sterilization. If necessary, parts can be
kept at 0°C or lower during the irradiation process.

PROCESSING

All commercially available biodegradable polymers can be melt processed by


conventional means such as injection molding, compression molding, and extrusion. As
with packaging, special consideration needs to be given to the exclusion of moisture from
the material before melt processing to prevent hydrolytic degradation. Special care must
be taken to dry the polymers before processing and to rigorously exclude humidity during
processing.

Because most biodegradable polymers have been synthesized by ring-opening


polymerization, a thermodynamic equilibrium exists between the forward or
polymerization reaction and the reverse reaction that results in monomer formation.
Excessively high processing temperatures may result in monomer formation during the
molding or extrusion process. The presence of excess monomer can act as a plasticizer,
changing the material's mechanical properties, and can catalyze the hydrolysis of the
device, thus altering degradation kinetics. Therefore, these materials should be processed
at the lowest temperatures possible.

Factors That Accelerate Polymer Degradation


• More hydrophilic backbone.
• More hydrophilic endgroups.
• More reactive hydrolytic groups in the backbone.
• Less crystallinity.
• More porosity.
• Smaller device size.

DEGRADATION

Once implanted, a biodegradable device should maintain its mechanical properties until it
is no longer needed and then be absorbed and excreted by the body, leaving no trace.
Simple chemical hydrolysis of the hydrolytically unstable backbone is the prevailing
mechanism for the polymer's degradation. This occurs in two phases. In the first phase,
water penetrates the bulk of the device, preferentially attacking the chemical bonds in the
amorphous phase and converting long polymer chains into shorter water-soluble
fragments. Because this occurs in the amorphous phase initially, there is a reduction in
molecular weight without a loss in physical properties, since the device matrix is still
held together by the crystalline regions. The reduction in molecular weight is soon
followed by a reduction in physical properties, as water begins to fragment the device
(see Figure 10). In the second phase, enzymatic attack and metabolization of the
fragments occurs, resulting in a rapid loss of polymer mass. This type of degradation—
when the rate at which water penetrates the device exceeds that at which the polymer is
converted into water-soluble materials (resulting in erosion throughout the device)—is
called bulk erosion. All of the commercially available synthetic devices and sutures
degrade by bulk erosion.

Figure 10. Generic absorption curves showing the


sequence of polymer molecular weight, strength,
and mass reduction. (Figure reproduced courtesy
of Journal of Craniofacial Surgery, (8)2:89, 1997.)

A second type of biodegradation, known as


surface erosion, occurs when the rate at which the
polymer penetrates the device is slower than the
rate of conversion of the polymer into water-
soluble materials. Surface erosion results in the device thinning over time while
maintaining its bulk integrity. Polyanhydrides and polyorthoesters are examples of
materials that undergo this type of erosion—when the polymer is hydrophobic, but the
chemical bonds are highly susceptible to hydrolysis. In general, this process is referred to
in the literature as bioerosion rather than biodegradation.

• The degradation-absorption mechanism is the result of many interrelated factors,


including:
• The chemical stability of the polymer backbone.
• The presence of catalysts, additives, impurities, or plasticizers.
• The geometry of the device.

Balancing these factors by tailoring an implant to slowly degrade and transfer stress at the
appropriate rate to surrounding tissues as they heal is one of the major challenges facing
researchers today.

COMMERCIAL BIODEGRADABLE DEVICES

The total U.S. revenues from commercial products developed from absorbable polymers
in 1995 was estimated to be over $300 million, with more than 95% of revenues
generated from the sale of bioabsorbable sutures. The other 5% is attributed to orthopedic
fixation devices in the forms of pins, rods, and tacks; staples for wound closure; and
dental applications.2 Research into biodegradable systems continues to increase, from the
60 to 70 papers published each year in the late 1970s to the more than 400 each year in
the early 1990s. The rate at which bioabsorbable fixation devices are cleared through the
FDA 510(k) regulatory process is also increasing, with seven devices cleared for sale in
1995.

What follows is a brief overview of some of the significant commercial applications of


biodegradable polymers.

Sutures. While comprising the lion's share of the total medical biodegradables market in
1995, this is a mature area not expected to grow rapidly in the future. About 125 million
synthetic bioabsorbable sutures are sold each year in the United States. They are divided
into braided and monofilament categories. Braided sutures are typically more pliable than
monofilament and exhibit better knot security when the same type of knot is used.
Monofilament sutures are more wiry and may require a more secure knot. Their major
advantage is that they exhibit less tissue drag, a characteristic that is especially important
for cardiovascular, ophthalmic, and neurological surgeries. A recent source in the
literature lists eight objective and three subjective parameters for suture selection based
on criteria such as tensile strength, strength retention, knot security, tissue drag, infection
potential, and ease of tying.3

Dental Devices. Biodegradable polymers have found use in two dental applications.
Employed as a void filler following tooth extraction, porous polymer particles can be
packed into the cavity to aid in quicker healing. As a guided-tissue-regeneration (GTR)
membrane, films of biodegradable polymer can be positioned to exclude epithelial
migration following periodontal surgery. The exclusion of epithelial cells allows the
supporting, slower-growing tissue—including connective and ligament cells—to
proliferate. Three examples of these GTR materials are Resolut from W.L. Gore
(Flagstaff, AZ), Atrisorb from Atrix Laboratories (Fort Collins, CO), and Vicryl mesh
from Ethicon.
Orthopedic Fixation Devices. Orthopedic fixation devices made from synthetic
biodegradable polymers have advantages over metal implants in that they transfer stress
over time to the damaged area, allowing healing of the tissues, and eliminate the need for
a subsequent operation for implant removal. The currently available materials have not
exhibited sufficient stiffness to be used as bone plates for support of long bones, such as
the femur. Rather, they have found applications where lower-strength materials are
sufficient: for example, as interference screws in the ankle, knee, and hand areas; as tacks
and pins for ligament attachment and meniscal repair; as suture anchors; and as rods and
pins for fracture fixation. Screws and plates of poly(l-lactide-co-glycolide) for
craniomaxillofacial repair have recently been cleared for marketing in the United States
under the trade name LactoSorb Craniomaxillofacial Fixation System (Biomet, Inc.,
Warsaw, IN).

Other Applications. Biodegradable polymers have found other applications that have
been commercialized or are under investigation. Anastomosis rings have been developed
as an alternative to suturing for intestinal resection. Tissue staples have also replaced
sutures in certain procedures. Other applications currently under scrutiny include ligating
clips, vascular grafts, stents, and tissue-engineering scaffolds. A list of commercial
synthetic biodegradable polymer devices by category is given in Table II.
Application Trade Name Composition a Manufacturer

Dexon PGA Davis and Geck

Maxon PGA-TMC Davis and Geck

Vicryl PGA-LPLA Ethicon

Sutures Monocryl PGA-PCL Ethicon

PDS PDO Ethicon Table II. Some commercial


biodegradable medical
Polysorb PGA-LPLA U.S. Surgical
products.
Biosyn PDO-PGA-TMC U.S. Surgical
Biodegradable Polymers
PGA Suture PGA Lukens
in Tissue Engineering

One of the exciting current


Sysorb DLPLA Synos areas for applications of
Endofix PGA-TMC or Acufex biodegradable polymers is
LPLA in tissue engineering.
Several companies are
Arthrex LPLA Arthrex investigating using these
Interference Bioscrew LPLA Linvatec materials as a matrix for
screws living cells. Important
properties in this regard
Phusiline LPLA-DLPLA Phusis include porosity for cell in-
Biologically PGA-DLPLA Instrument Makar growth, a surface that
Quiet balances hydrophilicity and
hydrophobicity for cellular
attachment, mechanical
Suture anchor Bio-Statak LPLA Zimmer properties that are
compatible with those of
Suretac PGA-TMC Acufex the tissue, and degradation
rate and by-products. The
polymer matrix may
Anastomosis Lactasorb LPLA Davis and Geck represent the device itself,
clip or can be a scaffold for cell
Anastomosis Valtrac PGA Davis and Geck growth in vitro that is
ring degraded by the growing
cells prior to implantation.
Dental Drilac DLPLA THM Biomedical The device can also be

Angioplastic Angioseal PGA-DLPLA AHP formulated to contain


plug additives or active agents
for more rapid tissue
Screw SmartScrew LPLA Bionx growth or compatibility.
For example, a bone
implant may contain a
Pins and rods Biofix LPLA or PGA Bionx form of calcium phosphate
or a growth factor such as
Resor-Pin LPLA-DLPLA Geistlich
one of the bone
morphogenetic proteins.
Tack SmartTack LPLA Bionx

Plates, mesh, LactoSorb PGA-LPLA Lorenz


screws

Antrisorb DLPLA Atrix


There are a number of ways of making the three-dimensional matrices required for tissue
engineering. These methods include woven or nonwoven preparations from spun fibers,
blown films using solvents or propellants, or sintered polymer particles. One of the
newest methods is being developed by Therics (Princeton, NJ), which has licensed a
system for building three-dimensional devices for use as scaffolds and in drug delivery
products. In this system, small spheres of polymer are laid out in thin films. Using
technology similar to that found in ink-jet printers, small amounts of solvent are used to
fuse particles together. The particles not fused are removed and another layer of particles
laid out. This particle placement and fusing is continued for many layers, until the exact
three-dimensional structure is obtained. Because each polymer layer is applied in a
separate step, different polymers can be used to obtain different properties in the interior
and exterior of the device.

Polymerization is the formation of long, repeating organic polymer chains.

Polymerization dates back to the beginning of DNA based life, as both DNA and proteins
can be considered polymers. The first 'synthetic' polymers of the 19th century were
actually formed by modifying natural polymers. For example nitrocellulose was
manufactured by reacting cellulose with nitric acid. The first genuinely man-made
polymer, bakelite, was synthesized in 1872, however research into polymers and
polymerization really accelerated in the 1930s after the serendipitous discovery of
polyethene by the chemical company ICI.

There are many forms of polymerization, and different systems exist to categorize them.
Categorizations include the addition-condensation system and the chain growth-step
growth system. Another form of polymerization is ring-opening polymerization, which is
similar to chain polymerization.

Addition polymerization involves the linking together of molecules incorporating double


or triple chemical bonds. These unsaturated monomers (the identical molecules which
make up the polymers) have extra, internal, bonds which are able to break and link up
with other monomers to form the repeating chain. Addition polymerization is involved in
the manufacture of polymers such as polyethene, polypropylene and polyvinylchloride
(PVC.) A special case of addition polymerization leads to living polymerization.

Condensation polymerization occurs when monomers bond together through


condensation reactions. Typically these reactions can be achieved through reacting
molecules incorporating alcohol, amine or carboxylic acid (or other carboxyl derivative)
functional groups. When an amine reacts with a carboxylic acid an amide or peptide bond
is formed, with the release of water (hence condensation polymerization.) This is the
process through which amino acids link up to form proteins, as well as how kevlar is
formed.

The chain growth-step growth system categorizes polymers based on their mechanism.
While most polymers will fall into their similar category from the addition-condensation
method of categorization, their are a fwe exceptions.
Chain growth polymers are defined as polymers formed by the reaction of monomer with
a reactive center. These polymers grow to high molecular weight at a very fast rate. It is
important to note that the overall conversion rates between chain and step growth
polymers are similar, but that high molecular weight polymers are formed in addition
reactions much more quickly than with step polymerizations.

Step growth polymers are defined as polymers formed by the stepwise reaction between
functional groups of monomer. Most step growth polymers are also classified as
condensation polymers, but not all step growth polymers (like polyurethanes formed from
isocyanate and alcohol bifunctional monomers) release condensates. Step growth
polymers increase in molecular weight at a very slow rate at lower conversions and only
reach moderately high molecular weights at very high conversion (i.e. >95%).

To alleviate inconsistencies in these naming methods, adjusted definitions for


condensation and addition polymers have been developed. A condensation polymer is
defined as a polymer that involves elimination of small molecules during its synthesis, or
contains functional groups as part of its backbone chain, or it repeat unit does not contain
all the atoms present in the hypothetical monomer to which it can be degraded.
Addition polymerization involves the breaking of double or triple bonds, which are used
to link monomers in to chains. In the polymerization of ethene (fig. 1), its double
bond is broken and it is used to bond to another poly(ethene) monomer. There are
several mechanisms through which this can be initiated. The free radical
mechanism was one of the first methods to be used. Free radicals are very reactive
atoms or molecules which have unpaired electrons. Taking the polymerization of
ethene as an example, the free radical mechanism can be divided in to three stages:
initiation, propagation and termination.

Ethene polymerization.png

Initiation is the creation of free radicals necessary for propagation. The radicals can be
created from organic peroxide molecules, molecules containing an O-O single bond, by
reacting oxygen with ethene. The products formed are unstable and easily break down
into two radicals. In an ethene monomer, one electron pair is held securely between the
two carbons in a sigma bond. The other is more loosely held in a pi bond. The free radical
uses one electron from the pi bond to form a more stable bond with the carbon atom. The
other electron returns to the second carbon atom, turning the whole molecule in to
another radical.

Propagation is the rapid reaction of this radicalised ethene molecule with another ethene
monomer, and the subsequent repetition to create the repeating chain.

Termination only occurs when two radical chains collide. The lone electrons pair up and
a stable molecule is formed, the product being a sum of the individual polymer chains.

Free radical addition polymerization must take place at high temperatures and pressures,
approximately 300°C and 2000 At. There are problems with the lack of control in the
reaction, specifically with the creation of variably branched chains. Also, as termination
occurs randomly, when two chains collide, it is impossible to control the length of
individual chains. Finally, reactions involving larger molecules such as polypropene are
difficult. For this reason new mechanisms for addition polymerization were developed.
An early replacement was the Ziegler-Natta catalyst.

The problem of branching occurs during propagation, when a chain curls back on itself
and breaks - leaving irregular chains sprouting from the main carbon backbone.
Branching makes the polymers less dense and results in low tensile strength and melting
points. Developed by Karl Ziegler and Giulio Natta in the 1950s, Ziegler-Natta catalysts
(triethylaluminium in the presence of a metal (IV) chloride) largely solved this problem.
Instead of a free radical reaction, the initial ethene monomer inserts between the
aluminium atom and one of the ethyl groups in the catalyst. The polymer is then able to
grow out from the aluminium atom and results in almost totally unbranched chains. With
the new catalysts, the tacticity of the polypropene chain, the alignment of alkyl groups,
was also able to be controlled. Different metal chlorides allowed the selective production
of each form i.e., syndiotactic, isotactic and atactic polymer chains could be selectively
created.

However there were further complications to be solved. If the Ziegler-Natta catalyst was
poisoned or damaged then the chain stopped growing. Also, Ziegler-Natta monomers
could be only small, and it was impossible to control the molecular mass of the polymer
chains. Again new catalysts, the metallocenes, were developed to tackle these problems.
Due to their structure they have less premature chain termination and branching.

Monomers: The Double Bond

In order for polymerization to occur with vinyl monomers, the substituents on the double
bond must be able to stabilize a negative charge. Stabilization occurs through
delocalization of the negative charge. Because of the nature of the carbanion propagating
center, substituents that react with bases or nucleophiles either must not be present or be
protected.

Vinyl monomers with substituents that stabilize the negative charge through charge
delocalization, undergo polymerization without termination or chain transfer. These
monomers include styrene, dienes, methacrylate, vinyl pyridine,aldehydes, epoxide,
episulfide, cyclic siloxane, and lactones. Polar monomers, using controlled conditions
and low temperatures, can undergo anionic polymerization. However, at higher
temperatures they do not produce living stable, carbanionic chain ends because their
polar substituents can undergo side reactions with both initiators and propagating chain
centers. The effects of counterion, solvent, temperature, Lew base additives, and
inorganic solvents have been investigated to increase the potential of anionic
polymerizations of polar monomers. Polar monomers include acrylonitrile,
cyanoacrylate, propylene oxide, vinyl ketone, acrolein, vinyl sulfone, vinyl sulfoxide,
vinyl silane and isocyanate.
You know from your studies that a double bond is two pairs of electrons being shared by
two atoms. This arrangement is fairly strong, but when other molecules - molecules that
like to react with electrons - are near, one of these pairs of electrons is vulnerable to
attack. One such attacking species is the free radical.

Below is an example of how a free radical forms (this one starts with peroxide, HOOH):
The peroxide molecule has an easy-to-break O-O single bond. Heat or light energy can
break this O-O single bond ().

The single dot represents one electron from the O-O peroxide bond. The HO· fragments
are the free radicals, and they are very unstable and reactive.
Free radicals are very reactive. When a free radical gets close to a double bond, one of
the bonds is disrupted. One of the electrons in the double bond is attracted to the free
radical. The double bond breaks, and a new single bond is formed ( ).

Notice that in forming this bond, one electron from the double bond is left alone. Thus,
another (larger) free radical has been formed.

By the way, we say that the free radical adds to the double bond. Get it? Addition
polimarization.

Let's look at the steps involved in a typical addition polymerization.

The Mechanism of Addition Polymerization

The formation of a polymer by addition polymerization is an example of a chain reaction.


Once a chain reaction gets started, it is able to keep itself going. The three steps of this
reaction to focus on are
how the reaction gets started (INITIATION)
how the reaction keeps going (PROPAGATION)
how the reaction stops (TERMINATION)

A Note About This Example:


There are various methods used to carry out addition polymerization chain reactions. The
details vary according to the method used.

We will focus on a commonly used mechanism involving a free radical. Our example
polymerization will combine ethylene (ethene) monomers (CH2=CH2), so our product
will be polyethylene. (Polyethylene is used to make food wrap, milk jugs, garbage bags,
and many other plastic products.)
I. INITIATION

The reactivity of initiators used in anionic polymerization should be similar to that of the
monomer that is the propagating species. The pKa values for the conjugate acids of the
carbanions formed from monomers can be used to deduce the reactivity of the monomer.
The least reactive monomers have the largest pKa values for their corresponding
conjugate acid and thus, require the most reactive initiator. Two main initiation pathways
involve electron transfer (through alkali metals) and strong anions.

If you looked at Part 3 of this tutorial, you have already seen the first part of the initiation
step of addition polimarization chain reaction. A peroxide molecule breaks up into two
reactive free radicals. Light or heat can provide the energy needed for this process.

We can write an equation for this process:

The second part of initiation occurs when the free radical initiator attacks and attaches to
a monomer molecule. This forms a new free radical, which is called the activated
monomer.

We can write an equation for this process, too:

Initiation by Electron Transfer

Szwarc and coworkers studied the initiation of polymerization through the use of
aromatic radical-anions such as sodium naphthenate. In this reaction, an electron is
transferred from the alkali metal to naphthalene. Polar solvents are necessary for this type
of initiation both for stability of the anion-radical and to solvate the cation species
formed.The anion-radical can then transfer an electron to the monomer.
Initation through electron transfer.

Initiation can also involve the transfer of an electron from the alkali metal to the
monomer to form an anion-radical. Initiation occurs on the surface of the metal, with the
reversible transfer of an electron to the adsorbed monomer.

Initiation by Strong Anions

Nucleophilic initiators include covalent or ionic metal amides, alkoxides, hydroxides,


cyanides, phosphines, amines and organometallic compounds (alkyllithium compounds
and Grignard reagents). The initiation process involves the addition of a neutral (B:) or
negative (B:-) nucleophile to the monomer

Initiation through strong anion.

The most commercially useful of these initiators has been the alkyllithium initiators.
They are primarily used for the polymerization of styrenes and dienes

II. PROPAGATION

Propagation in anionic addition polymerization results in the complete consumption of


monomer. It is very fast and occurs at low temperatures. This is due to the anion not
being very stable, the speed of the reaction as well as that heat is released during the
reaction. The stability can be greatly enhanced by reducing the temperatures to near 0˚C.
The propagation rates are generally fairly high compared to the decay reaction, so the
overall polymerization rates is generally not affected.
Propagation of an anionic addition polymerization

During a chain reaction, most of the time is spent in the propagation phase as the polymer
chain grows. In the propagation phase, the newly-formed activated monomer attacks and
attaches to the double bond of another monomer molecule. This addition occurs again
and again to make the long polymer chain.

Once again, we can write an equation for this reaction:

The "n" stands for any number of monomer molecules, typically in the thousands.
III. TERMINATION

Anionic addition polymerizations have no formal termination pathways because proton


transfer from solvent or other positive species does not occur. However, termination can
occur through unintentional quenching due to trace impurities. This includes trace
amounts of oxygen, carbon dioxide or water. Intentional termination can occur through
the addition of water or alcohol. Another method of termination, chain transfer, can occur
when an agent can act as a Bronsted acid .In this case, the pKa value of the agent is
similar to the conjugate acid of the propagating carbanionic chain end. Spontaneous
termination occurs because the concentration of carbanion centers decay over time and
eventually results in hydride elimination. Polar monomers are more reactive because they
are stabilized by their polar substituents. These polar substituents can react with
nucleophiles which results in termination as well as side reactions that compete with both
initation and propagation.

This chain reaction cannot go on forever. The reaction must terminate, but how? A
growing polymer chain joins with another free radical. We watched a peroxide break up
to form two radicals. It makes sense that two free radicals could join to make a stable
bond.

The equation representing this step of the chain reaction can be written simply as:

Remember: The R and R' groups here can be the original free radicals, the growing
polymer chains, or even one of each. Termination reactions can, however, be more
complicated looking.

An Important Note:
Chemists can control the way a polymer does each of these steps by varying the reactants,
the reaction times, and the reaction conditions.

The physical properties of a polymer chain depend on the polymer's average length, the
amount of branching, and the constituent monomers.

This is an exciting and useful field of chemistry!

A Simulation of Addition Polymerization

In Part 4 of this tutorial, you saw that there are three steps in an addition polymerization
chain reaction. You also saw that there are only two kinds of molecules in the chain
reaction: the initiator molecule and the monomers. Polymerization begins at the initiator,
and reaction continues until there are no more monomers to add to the growing polymer
chain. The chain grows only at the reactive end, the end with the unpaired electron.
The simulation you will see displays this process graphically. Click on the button below
to view the Simulation window. (If nothing happens, click here.)

The larger box in this window is the area where you'll see the polymerization of
monomers, represented by black balls. The initiator molecules are in red. You input the
number of initiators, press START, and the monomers will add onto the initiators
linearly. As the average chain length increases, you'll see it displayed graphically in the
smaller box. The graph is a bar graph of the average size of the polymer chain versus the
reciprocal of the number of initiators (a red bar represents the active or most recent
polymerization; a blue bar, the past polymerizations).

Note:
You may not see the full length of the chain, but the numbers you see in and below the
graph are correct for the given number of initiators.

Some Assumptions:
First, we assume that the red initiator molecule is the activated monomer that you saw in
Part 4. Second, we assume that there are only 200 monomers in the polymerization. In
real life, the number of monomers are on the order of 1023. Despite the low number of
monomers in the simulation, it does show the correct, real-life trend of how the number
of initiators affects the average chain length. Third, polymerization is terminated when
the monomers run out. There is no visual coupling of free radicals; there are as many
polymers as there are initiator molecules.

View the simulation several times with different numbers of initiators to see a trend in the
bar graph. The more initiators, the shorter the chains (if there is a constant number of
monomers).

Addition polymerization is not the only mechanism by which polymerization can occur.
Coordination polymerization is a form of addition polymerization in which monomer
adds to a growing macromolecule through an organometallic active center. The
development of this polymerization technique started in the 1950s with heterogeneous
Ziegler-Natta catalysts based on titanium tetrachloride and an aluminium co-catalyst such
as methylaluminoxane. Coordination polymerization has a great impact on the physical
properties of vinyl polymers such as polyethylene and polypropylene compared to the
same polymers prepared by other techniques such as free radical polymerization. The
polymers tend to be linear and not branched and have much higher molar mass.
Coordination type polymers are also stereoregular and can be isotactic or syndiotactic
instead of just atactic. This tacticity introduces crystallinity in otherwise amorphous
polymers. From these differences in polymerization type the distinction originates
between low density polyethylene (LDPE), high density polyethylene (HDPE) or even
ultra high molecular weight polyethylene (UHMWPE). Polymerizations catalysed by
metallocenes occur via the Cossee-Arlman mechanism.

In many applications Ziegler-Natta polymerization is succeeded by metallocene catalysis


polymerization. This method is based on homogeneous metallocene catalysts such as the
Kaminsky catalyst discovered in the 1970s. The 1990s brought forward a new range of
post-metallocene catalysts.

Insertion of aluminum alkyls into olefins was studied by Ziegler:


Important discovery: R3Al + Lewis acids:

Another important discovery: tacticity control:

Results:

• Nobel Prize in Chemistry for Zeigler and Natta (1963)


• Multibillion $ industry

A typical Ziegler-Natta catalyst can be produced by mixing solutions of titanium(IV)


chloride (TiCl4) and triethylaluminum [Al(CH2CH3)3] dissolved in a hydrocarbon solvent
from which both oxygen and water have been rigorously excluded. The product of this
reaction is an insoluble olive-colored complex in which the titanium has been reduced to
the Ti(III) oxidation state.

The catalyst formed in this reaction can be described as coordinately unsaturated because
there is an open coordination site on the titanium atom. This allows an alkene to act as a
Lewis base toward the titanium atom, donating a pair of electrons to form a transition-
metal complex.

The alkene is then inserted into a Ti-CH2CH3 bond to form a growing polymer chain and
a site at which another alkene can bond.

Thus, the titanium atom provides a template on which a linear polymer with carefully
controlled stereochemistry can grow
Overall Scheme of Coordination Polymerization

• Limited to ethylene and other a-olefins like propylene. (Actually, it is the only
good way to polymerize these monomers.)
• Produces linear polymer, with very few branches (e.g., high density polyethylene,
HDPE).
• Capable of producing homo-tactic polymers.
• Most commercial initiators are insoluble complexes or supported on insoluble
carriers.
• Very complex mechanism, still poorly understood for the heterogeneous systems.
• Termination is almost exclusively by chain transfer.
• Modern "high mileage" initiators produce up to 1000's of kg per g initiator.
• Initiators are often called "catalysts" even though they are consumed by the
process. Many chains are started per molecule of initiator.
Mechanism of Coordination Polymerization

The mechanism is poorly understood because it takes place on the surface of an insoluble
particle, a difficult situation to probe experimentally. The mechanism shown below is one
of several models proposed to at least partially explain the action of the Ziegler-Natta
systems, but it is only an approximation of the more complex process that actually
occurs.
Hyperbranched polymers by coordination polymerization

Hyperbranched copolymers comprising at least one C2 -C20 α-monoolefin monomers and


0.2 to 20 mole % of at least one α,ω-non-conjugated diene monomers having 5 to 18
carbon atoms are prepared by coordination (metallocene) copolymerization of the
monomers and quenching the reaction prior to the formation of a gelled product. The
building blocks of the products are characterized by a number average molecular weight
less than 5 times the entanglement molecular weight of a homopolymer prepared using
the same catalyst but in the absence of the diene component.
Condensation polymers are any kind of polymers formed through a condensation
reaction, releasing small molecules as by-products such as water or methanol, as opposed
to addition polymers which involve the reaction of unsaturated monomers. Types of
condensation polymers include polyamides, polyacetals and polyesters.

Condensation polymerization, a form of step-growth polymerization, is a process by


which two molecules join together, resulting loss of small molecules which is often
water. The type of end product resulting from a condensation polymerization is
dependent on the number of functional end groups of the monomer which can react.

Monomers with only one reactive group terminate a growing chain, and thus give end
products with a lower molecular weight. Linear polymers are created using monomers
with two reactive end groups and monomers with more than two end groups give three
dimensional polymers which are crosslinked.

Dehydration synthesis often involves joining monomers with an -OH (hydroxyl) group
and a freely ionized -H on either end (such as a hydrogen from the -NH2 in nylon or
proteins). Normally, two or more different monomers are used in the reaction. The bonds
between the hydroxyl group, the hydrogen atom and their respective atoms break forming
water from the hydroxyl and hydrogen, and the polymer.

Polyester is created through ester linkages between monomers, which involve the
functional groups carboxyl and hydroxyl (an organic acid and an alcohol monomer).

Nylon is another common condensation polymer. It can be manufactured by reacting di-


amines with carboxyl derivatives. In this example the derivative is a di-carboxylic acid,
but di-acyl chlorides are also used. Another approach used is the reaction of di-functional
monomers, with one amine and one carboxylic acid group on the same molecule:

The carboxylic acids and amines link to form peptide bonds, also known as amide
groups. Proteins are condensation polymers made from amino acid monomers.
Carbohydrates are also condensation polymers made from sugar monomers such as
glucose and galactose.

Condensation polymerization is occasionally used to form simple hydrocarbons. This


method, however, is expensive and inefficient, so the addition polymer of ethene
(polyethylene) is generally used.

Condensation polymers, unlike addition polymers, may be biodegradable. The peptide or


ester bonds between monomers can be hydrolysed by acid catalysts or bacterial enzymes
breaking the polymer chain into smaller pieces.The most commonly known condensation
polymers are proteins, fabrics such as nylon, silk, or polyester.
Condensation
Condensation is an organic reaction when two molecules combine, usually in the
presence of a catalyst, with the elimination of water or some other simple molecule.
Catalysts commonly used in condensation reactions include acids and bases. The
combination of two identical molecules is known as self-condensation. This process
forms larger molecules, many of which are useful in organic synthesis. Aldehydes,
ketones, esters, alkynes, and alcohols are among several organic compounds that combine
with each other to form larger molecules.

Example:

CH3OH + CH3OH => CH3OCH3 + HOH

methanol + methanol => methoxymethane + water

For a carboxylic acid undergoing condensation reaction, it combines with another


reactant, forming two products - an organic compound and the byproduct of water. In an
event when a carboxylic acid reacts with an alochol to produce an ester and water, this
process is called esterification.

Example:

CH3COOH + HOCH3 => CH3COOCH3 + HOH

ethanoic acid + methanol => methyl ethanoate +


water

Polymerization is a process in which very small molecules, called monomers, combine


chemically with each other to produce a very large chainlike molecule, called a polymer.
The monomer molecules may be all alike, or they may represent two, three, or more
different compounds. The monomers react to form a polymer without the formation of
by-products such as water. The structure has one structural unit, or monomer, that occurs
repeatedly. Through polymerization of ethylene (ethene), CH2CH2, the structure of the
polymer can therefore be represented by -(CH2CH2)n- Where n can be several thousand.
Because of this, polymers have incredible molar masses up to millions of grams per mole.
Example:

H H H H H H H H H H H H
| | | | | | | | | | | |
C=C + C=C + C=C => :C-C:C-C:C-C: OR -(CH2CH2)n-
| | | | | | | | | | | |
H H H H H H H H H H H H

ethylene part of polyethylene

In condensation polymerization, two functional groups of two different monomer


molecules are joined together which produces a small molecule such as water. The
monomers bond at where the hydrogen atoms were taken out to produce water. In order
to become a condensed polymer, the monomer molecules must have at least two
functional groups. The combination of two identical molecules is known as self-
condensation. The reaction between a carboxylic acid and an alcohol creates an ester. If
the carboxylic acid and the alcohol were the monomers of the polymer, during
polymerization, they would create polyester, and produce water. The polymerization of a
carboxylic acid and an amine similarly creates polyamides.

Example:

COOHC6H6COOH + HOCH2CH2OH => COOHC6H6OCH2CH2OH + HOH

1,4-benzenedioic acid + 1,2-ethandiol =>


polyester + water

Condensation polymerization (part 1 of 2)

The bonds between monomer molecules are formed with the elimination of a small
molecule, such as water. This is due to one of the monomers losing a hydrogen atom, and
another losing a hydroxyl group. Each monomer usually has two functional groups of
condensation polymers, the polyesters, and the polyamides. The condensation reaction:
(diag.) Now consider some examples of polyesters; Polyethylene terephthalate (PET) is
produced by the step-growth polymerization of ethylene glycol and terephthalic acid.

The presence of the large benzene rings in the repeating units, gives the polymer notable
stiffness and strength, especially when the polymer chains are aligned with one another in
an orderly arrangement by stretching. In this semi-crystalline form, PET is made into a
high-strength textile fiber, where its stiffness makes them highly resistant to deformation,
so that they have excellent resistance to wrinkling in fabrics.
Monomers: Functional Groups

The monomers that are involved in condensation polymerization are not the same as
those in addition polymerization. The monomers for condensation polymerization have
two main characteristics:.
Instead of double bonds, these monomers have functional groups (like alcohol, amine,
or carboxylic acid groups).
Each monomer has at least two reactive sites, which usually means two functional
groups.

Some monomers have more than two reactive sites, allowing for branching between
chains, as well as increasing the molecular mass of the polymer. Four examples of these
difunctional monomers were introduced in Part 2 of this tutorial. Here they are again:

Guess the names of each of these monomers. Give the letter that
corresponds to the correct name of the structure (use each letter only
once). Hints: Glycol means that a molecule has more than one
alcohol (-OH) group. Amine means that a molecule has an amino (-
NH2) group. Diamine (or diamino) means that a molecule contains
two amino groups. Acid means that a molecule contains a carboxylic
acid group (-COOH). Click the button when done.

Let's look again at the functional groups on these monomers. We've seen three:

The carboxylic acid group


The amino group
The alcohol group

You might have learned in chemistry or biology class that these groups can combine in
such a way that a small molecule (often H2O) is given off.

The Amide Linkage:


When a carboxylic acid and an amine react, a water molecule is removed, and an amide
molecule is formed.

Because of this amide formation, this bond is known as an amide linkage.

The Ester Linkage:


When a carboxylic acid and an alcohol react, a water molecule is removed, and an ester
molecule is formed.

Because of this ester formation, this bond is known as an ester linkage.

In Summary:
Monomers involved in condensation polymerization have functional groups. These
functional groups combine to form amide and ester linkages. When this occurs, a water
molecule in removed. Since water is removed, we call these reactions condensation
reactions (water condenses out.). When a condensation reaction involves polymerization,
we call it condensation polimarization.

Let's look at a few common examples of condensation polymers.

The Mechanism of Condensation Polymerization

You know that monomers that are joined by condensation polymerization have two
functional groups. You also know (from Part 6) that a carboxylic acid and an amine can
form an amide linkage, jand a carboxylic acid and an alcohol can form an ester linkage.
Since each monomer has two reactive sites, they can form long-chain polymers by
making many amide or ester links. Let's look at two examples of common polymers made
from the monomers we have studied.
Example 1:
A carboxylic acid monomer and an amine monomer can join in an amide linkage.

As before, a water molecule is removed, and an amide linkage is formed. Notice that an
acid group remains on one end of the chain, which can react with another amine
monomer. Similarly, an amine group remains on the other end of the chain, which can
react with another acid monomer.

Thus, monomers can continue to join by amide linkages to form a long chain. Because of
the type of bond that links the monomers, this polymer is called a polyamide. The
polymer made from these two six-carbon monomers is known as nylon-6,6. (Nylon
products include hosiery, parachutes, and ropes.)

Example 2:
A carboxylic acid monomer and an alcohol monomer can join in an ester linkage.

A water molecule is removed as the ester linkage is formed. Notice the acid and the
alcohol groups that are still available for bonding. ( )

Because the monomers above are all joined by ester linkages, the polymer chain is a
polyester. This one is called PET, which stands for poly(ethylene terephthalate). (PET is
used to make soft-drink bottles, magnetic tape, and many other plastic products.)

Let's summarize:
As difunctional monomers join with amide and ester linkages, polyamides and polyesters
are formed, respectively. We have seen the formation of the polyamide nylon-6,6 and the
polyester PET. There are numerous other examples.

Remember: The above process is called condensation polymerization because a molecule


is removed during the joining of the monomers. This molecule is frequently water.
A Simulation of Condensation Polymerization

You learned in Part 7 that condensation polymers are made from monomers that have at
least two functional groups. Because of this, the polymers can grow at either end of the
chain.

During the polymerization process, the monomers tend to form dimers (two linked
monomers) and trimers (three linked monomers) first. Then, these very short chains react
with each other and with monomers. The overall result is that, at the beginning of
polymerization, there are many relatively short chains. It is only near the end of
polymerization that very long chains are formed.

The simulation you will see displays this process graphically. Click on the button below
to view the Simulation window. (If nothing happens, click here.)

The larger box in this window is the area where you'll see the polymerization of
monomers, represented by gray balls. When you click START, you'll see a lattice of
gray, or unreacted, monomers. Once they polymerize into dimers, trimers, and so on, the
monomers will turn black. Polymerization will continue for a few seconds. Then the
display will change into a bar graph entitled "Distribution" and show the progression of
the polymerization over time. The x-axis is the number of units in the polymer (the "n" in
the formula of a polymer). This is suggested graphically with the series of polymers
projected into the screen. As you move to the left, the polymers are longer. The y-axis is
the number of polymers. The higher the bar, the more numerous are the polymers. The
graph shows dynamically the distribution of polymers in the polymerization as the
reaction progresses. Notice that at the beginning of the polymerization, the distribution
lies farther to the right, meaning that there are a lot of monomers, dimers, trimers, and
other short chains but few long chains. As the polymerization progresses, the distribution
shifts to the left, indicating that there are fewer short chains and more of the longer ones.

The smaller box is a graph that displays the average size of the polymers versus the time
of the polymerization. Again, notice that at early times, there are mostly short chains, and
that near the end, there are more long.

Some Assumptions:
First, we assume that there is only one type of difunctional monomer, as opposed to two
types, as you saw in the two examples in Part 7. If you imagine that the polymers in the
simulation are polyamides (like nylon-6,6), then the monomer has one carboxylic acid
group and one alcohol group (picture the dimer you saw in Example 1 in the previous
section). Second, we assume that there are only 90,000 monomers in the polymerization.
In real life, the number of monomers is on the order of 1023. Despite the low number of
monomers in the simulation, it does show the correct, real-life distribution of polymer
chains over time
Step-growth polymerization
Step growth polymerization, also called condensation polymerization, refers to
polymerizations in which bi-functional or multifunctional monomers react to form
dimers, trimers, longer oligomers and eventually long chain polymers. Many naturally
occurring and some synthetic polymers are produced by step-growth polymerization. It
requires a high extent of reaction to achieve high molecular weight, thus only a few kinds
of commercial polymers can be synthesized this way, like polyester, polyamide,
polyurethane,etc. The easiest way to visualize a step growth polymerization is a group of
people holding hands to form a human chain: each person has two hands (=reactive sites).

A monomer with functionality 3 will introduce branching in a polymer and will


ultimately form a cross-linked macrostructure or network even at low fractional
conversion. The point at which this three-dimensional structure is formed is known as the
gel point because it is signalled by an abrupt change in viscosity. One of the earliest so-
called thermosets is known as bakelite. It is not always water that is released in step-
growth polymerization: in acyclic diene metathesis or ADMET dienes polymerize with
loss of ethylene.

A generic representation of a step-growth polymerization. (Single white dots represent


monomers and black chains represent oligomers and polymers)
Comparison of Molecular weight vs conversion plot between step-growth and chain-
growth polymerization

This technique is usually compared with chain-growth polymerization to show its


characteristics.

Step-growth polymerization Chain-growth polymerization

Growth by addition of monomer only at one end of


Growth throughout matrix
chain

Rapid loss of monomer early in the reaction Some monomer remains even at long reaction times

Different mechanisms operate at different stages of


Same mechanism throughout
reaction (i.e. Initiation, propagation and termination)

Average molecular weight increases slowly at low Molar mass of backbone chain increases rapidly at
conversion and high extents of reaction are required early stage and remains approximately the same
to obtain high chain length throughout the polymerization

Ends remain active (no termination) Chains not active after termination

No initiator necessary Initiator required


Classes of step-growth polymers

Polyester has high Tg, high Tm, good mechanical properties to about 175°C, good
resistance to solvent and chemicals. It can exist as fibers and films. The former is used in
garments, felts, tire cord, etc. The latter appears in magnetic recording tape and high
grade films.

Polyamide (nylon) has good balance of properties: high strength, good elasticity and
abrasion resistance, good toughness, favorable solvent resistance. The applications of
polyamide include: rope, belting, fiber cloths, thread, substitute for metal in bearings,
jackets on electrical wire.

Polyurethane can exist as elastomers with good abrasion resistance, hardness, good
resistance to grease and good elasticity, as fibers with excellent rebound, as coatings with
good resistance to solvent attack and abrasion and as foams with good strength, good
rebound and high impact strength.

Polyurea shows high Tg, fair resistance to greases, oils and solvents. It can be used in
truck bed liners, bridge coating, caulk and decorative designs.

Polysiloxane are available in a wide range of physical states-from liquids to greases,


waxes, resins and rubbers. Uses of this material are as antifoam and release agents,
gaskets, seals, cable and wire insulation, hot liquids and gas conduits, etc.

Polycarbonates are transparent, self-extinguishing materials. They possess properties


like crystalline thermoplasticity, high impact strength, good thermal and oxidative
stability. They can be used in machinery, auto-industry and medical applications. For
example, the cockpit canopy of F-22 Raptor is made of high optical quality
polycarbonate.

Polysulfides have outstanding oil and solvent resistance, good gas impermeability, good
resistance to aging and ozone. However, it smells bad and it shows low tensile strength as
well as poor heat resistance. It can be used in gasoline hoses, gaskets and places that
require solvent resistance and gas resistance.

Polyether shows good thermoplastic behavior, water solubility, generally good


mechanical properties, moderate strength and stiffness. It is applied in sizing for cotton
and synthetic fibers, stabilizers for adhesives, binders, and film formers in
pharmaceuticals.

Phenol formaldehyde resin (Bakelite) have good heat resistance, dimensional stability
as well as good resistance to most solvents. It also shows good dielectric properties. This
material is typically used in molding applications, electrical, radio, televisions and
automotive parts where their good dielectric properties are of use. Some other uses
include: impregnating paper, varnishes, decorative laminates for wall coverings.
Advances in step-growth Polymers
The driving force in designing new polymers is the prospect of replacing other materials
of construction especially metals using lightweight and heat-resistant polymers. The
advantages of lightweight polymers include: high strength, solvent and chemical
resistance, contributing to a variety of potential uses, such as electrical and engine parts
on automotive and aircraft components, coatings on cookware, coating and circuit boards
for electronic and microelectronic devices, etc. Polymer chains based on aromatic rings
are desirable due to high bond strengths and rigid polymer chains. High molecular weight
and crosslinking are desirable for the same reason. Strong dipole-dipole, hydrogen bond
interactions and crystallinity also improve heat resistance. To obtain desired mechanical
strength, sufficiently high molecular weights are necessary, however, decreased solubility
is a problem. One approach to solve this problem is to introduce of some flexibilizing
linkages such as isopropylidene, C=O, and SO2 into the rigid polymer chain by using an
appropriate monomer or comonomer. Another approach involves the synthesis of reactive
telechelic oligomers containing functional end groups capable of reacting with each
other, polymerization of the oligomer gives higher molecular weight, referred to as chain
extension.

Aromatic polyether

The oxidative coupling polymerization of many 2,6-disubstituted phenols using a


catalytic complex of a cuprous salt and amine form aromatic polyethers, commercially
referred to as poly(p-phenylene oxide) or PPO. Neat PPO has little commercial uses due
to its high melt viscosity. Its available products are blends of PPO with high-impact
polystyrene (HIPS).
Polyethersulfone

Polyethersulfone (PES) is also referred to as polyetherketone, polysulfone. It is


synthesized by nucleophilic aromatic substitution between aromatic dihalides and
bisphenolate salts. Polyethersulfones are partially crystalline, highly resistant to a wide
range of aqueous and organic environment. They are rated for continuous service at
temperatures of 240-280 oC. The polyketones are finding applications in areas like
automotive, aerospace, electrical-electronic cable insulation.

Aromatic polysulfides

Poly(p-phenylene sulfide) (PPS) is synthesized by the reaction of sodium sulfide with


[[p-dichlorobenzene]] in a polar solvent such as 1-methyl-2-pyrrolidinone (NMP). It is
inherently flame resistant and stable toward organic and aqueous conditions, however, it
is somewhat susceptible to oxidants. Applications of PPS include automotive, microwave
oven component, coating for cookware when blend with fluorocarbon polymers and
protective coatings for valves, pipe, electromotive cells etc.

Aromatic polyimide
Aromatic polyimide are synthesized by the reaction of dianhydrides with diamines, for
example, pyromellitic anhydride with [[p-phenylenediamine]]. It can also be
accomplished using diisocyanates in place of diamines. Solubility considerations
sometimes result in using the half acid-half ester of the dianhydride instead of the
dianhydride. Polymerization is accomplished by a two-stage process due to insolubility of
polyimdes. The first stage forms a soluble and fusible high-molecular-weight poly(amic
acid) in a polar aprotic solvent such as NMP or N,N-dimethylacetamide. The poly(amic
aicd) can then be processed into the desired physical form of the final plymer product
(e.g., film, fiber, laminate, coating) which is insoluble and infusible.

Telechelic oligomer approach

Telechelic oligomer approach applies the usual polymerization manner except that one
includes a monofunctional reactant to stop reaction at the oligomer stage, generally in the
50-3000 molecular weight. The monofunctional reactant not only limits polymerization
but end-caps the oligomer with functional groups capable of subsequent reaction to
achieve curing of the oligomer. Functional groups like alkyne, norbornene, maleimide,
nitrite, and cyanate have been used for this purpose. Maleimide and norbornene end-
capped oligomers can be cured by heating. Alkyne, nitrile and cyanate end-capped
oligomers can undergo cyclotrimerization yielding aromatic structures.
Thermoplastic
A thermoplastic is a polymer that turns to a liquid when heated and freezes to a very
glassy state when cooled sufficiently. Most thermoplastics are high-molecular-weight
polymers whose chains associate through weak Van der Waals forces (polyethylene);
stronger dipole-dipole interactions and hydrogen bonding (nylon); or even stacking of
aromatic rings (polystyrene). Thermoplastic polymers differ from thermosetting polymers
(Bakelite; vulcanized rubber) as they can, unlike thermosetting polymers, be remelted
and remoulded. Many thermoplastic materials are addition polymers; e.g., vinyl chain-
growth polymers such as polyethylene and polypropylene.

Stress strain graph of thermoplastic material.

Thermoplastics are elastic and flexible above a glass transition temperature Tg, specific
for each one — the midpoint of a temperature range in contrast to the sharp melting point
and melting point of a pure crystalline substance like water. Below a second, higher
melting temperature, Tm, also the midpoint of a range, most thermoplastics have
crystalline regions alternating with amorphous regions in which the chains approximate
random coils. The amorphous regions contribute elasticity and the crystalline regions
contribute strength and rigidity, as is also the case for non-thermoplastic fibrous proteins
such as silk. (Elasticity does not mean they are particularly stretchy; e.g., nylon rope and
fishing line.) Above Tm all crystalline structure disappears and the chains become
randomly inter dispersed. As the temperature increases above Tm, viscosity gradually
decreases without any distinct phase change.

Thermoplastics can go through melting/freezing cycles repeatedly and the fact that they
can be reshaped upon reheating gives them their name. This quality makes thermoplastics
recyclable. The processes required for recycling vary with the thermoplastic. The plastics
used for soda bottles are a common example of thermoplastics that can be and are widely
recycled. Animal horn, made of the protein α-keratin, softens on heating, is somewhat
reshapable, and may be regarded as a natural, quasi-thermoplastic material.
Some thermoplastics normally do not crystallize: they are termed "amorphous" plastics
and are useful at temperatures below the Tg. They are frequently used in applications
where clarity is important. Some typical examples of amorphous thermoplastics are
PMMA, PS and PC. Generally, amorphous thermoplastics are less chemically resistant
and can be subject to stress cracking. Thermoplastics will crystallize to a certain extent
and are called "semi-crystalline" for this reason. Typical semi-crystalline thermoplastics
are PE, PP, PBT and PET. The speed and extent to which crystallization can occur
depends in part on the flexibility of the polymer chain. Semi-crystalline thermoplastics
are more resistant to solvents and other chemicals. If the crystallites are larger than the
wavelength of light, the thermoplastic is hazy or opaque. Semi-crystalline thermoplastics
become less brittle above Tg. If a plastic with otherwise desirable properties has too high
a Tg, it can often be lowered by adding a low-molecular-weight plasticizer to the melt
before forming (Plastics extrusion; molding) and cooling. A similar result can sometimes
be achieved by adding non-reactive side chains to the monomers before polymerization.
Both methods make the polymer chains stand off a bit from one another. Before the
introduction of plasticizers, plastic automobile parts often cracked in cold winter weather.
Another method of lowering Tg (or raising Tm) is to incorporate the original plastic into a
copolymer, as with graft copolymers of polystyrene, or into a composite material.
Lowering Tg is not the only way to reduce brittleness. Drawing (and similar processes
that stretch or orient the molecules) or increasing the length of the polymer chains also
decrease brittleness.

Although modestly vulcanized natural and synthetic rubbers are stretchy, they are
elastomeric thermosets, not thermoplastics. Each has its own Tg, and will crack and
shatter when cold enough so that the crosslinked polymer chains can no longer move
relative to one another. But they have no Tm and will decompose at high temperatures
rather than melt. Recently, thermoplastic elastomers have become available.

Terminology

The literature on thermoplastics is huge, and can be quite confusing, as the same
chemical can be available in many different forms (for example, at different molecular
weights), which might have quite different physical properties. The same chemical can be
referred to by many different tradenames, by different abbreviations; two chemical
compounds can share the same name; a good example of the latter is the word "Teflon"
which is used to refer to a specific polymer (PTFE); to related polymers such as PFA, and
generically to fluoropolymers.

Furthermore, over the last 30 years, there has been tremendous change in the plastics
industry, with many companies going out of business or merging into other companies.
Many production plants frequently changed hands or have been relocated to emerging
countries in Eastern Europe or Asia, with different trademarks.
Testing

Testing of thermoplastics can take various forms.

Tensile tests — ISO 527 -1/-2 and ASTM D 638 set out the standardized test methods.
These standards are technically equivalent. However they are not fully comparable
because of the difference in testing speeds. The modulus determination requires a high
accuracy of ± 1 micrometer for the dilatometer.

Flexural tests — 3-points flexural tests are among the most common and classic
methods for semi rigid and rigid plastics.

Pendulum impact tests — impact tests are used to measure the behavior of materials at
higher deformation speeds. Pendulum impact testers are used to determine the energy
required to break a standardized specimen by measuring the height to which the
pendulum hammer rises after impacting the test piece.

Table of thermoplastics
Polymer Melting point

Acrylonitrile butadiene styrene (ABS) -

Acrylic (PMMA) 130–140 °C

Celluloid -

Polycaprolactone (PCL) 62 °C

Polyethylene (PE) 105-130 °C

Polyphenylene oxide (PPO) -

Polyimide (PI) -

Polymethylpentene (PMP) -

Polylactic acid (PLA) 50-80 °C

Polyphenylene sulfide (PPS) -

Polyphthalamide (PPA) -

Polypropylene (PP) -

Polystyrene (PS) 240 °C

Polyurethane (PU) -

Polyvinyl acetate (PVA) -

Polyvinyl chloride (PVC) 80 °C


Polylactic acid
Polylactic acid or polylactide (PLA) is a biodegradable, thermoplastic, aliphatic
polyester derived from renewable resources, such as corn starch (in the U.S.) or
sugarcanes (rest of world). Although PLA has been known for more than a century, it has
only been of commercial interest in recent years, in light of its biodegradability.

Chemical and physical properties

Due to the chiral nature of lactic acid, several distinct forms of polylactide exist: poly-L-
lactide (PLLA) is the product resulting from polymerization of L,L-lactide (also known
as L-lactide). PLLA has a crystallinity of around 37%, a glass transition temperature
between 50-80 °C and a melting temperature between 173-178 °C.

Polylactic acid can be processed like most thermoplastics into fiber (for example using
conventional melt spinning processes) and film. The melting temperature of PLLA can be
increased 40-50 °C and its heat deflection temperature can be increased from
approximately 60°C to up to 190 °C by physically blending the polymer with PDLA
(poly-D-lactide). PDLA and PLLA form a highly regular stereocomplex with increased
crystallinity. The temperature stability is maximised when a 50:50 blend is used, but even
at lower concentrations of 3-10% of PDLA, there is still a substantial improvement. In
the latter case, PDLA acts as a nucleating agent, thereby increasing the crystallization
rate. Biodegradation of PDLA is slower than for PLA due to the higher crystallinity of
PDLA. PDLA has the useful property of being optically transparent.

Applications

Biodegradable plastic cups in use at an eatery.

Stereocomplex blends of PDLA and PLLA have a wide range of applications, such as
woven shirts (ironability), microwavable trays, hot-fill applications and even engineering
plastics (in this case, the stereocomplex is blended with a rubber-like polymer such as
ABS). Such blends also have good form-stability and visual transparency, making them
useful for low-end packaging applications. Progress in bio-technology has resulted in the
development of commercial production of the D(-) form, something that was not possible
until recently.
PLA is currently used in a number of biomedical applications, such as sutures, stents,
dialysis media and drug delivery devices. It is also being evaluated as a material for tissue
engineering. Because it is biodegradable, it can also be employed in the preparation of
bioplastic, useful for producing loose-fill packaging, compost bags, food packaging, and
disposable tableware. In the form of fibers and non-woven textiles, PLA also has many
potential uses, for example as upholstery, disposable garments, awnings, feminine
hygiene products, and nappies.

PLA has been used as the hydrophobic block of amphiphilic synthetic block copolymers
used to form the vesicle membrane of polymersomes.

PLA is a sustainable alternative to petrochemical-derived products, since the lactides


from which it is ultimately produced can be derived from the fermentation of agricultural
by-products such as corn starch[1] or other carbohydrate-rich substances like maize, sugar
or wheat.

PLA is more expensive than many petroleum-derived commodity plastics, but its price
has been falling as production increases. The demand for corn is growing, both due to the
use of corn for bioethanol and for corn-dependent commodities, including PLA.

PLA has also been developed in the United Kingdom to serve as sandwich packaging

Thermosetting polymer
Thermosetting plastics (thermosets) are polymer materials that irreversibly cure. The
cure may be done through heat (generally above 200 degrees Celsius), through a
chemical reaction (two-part epoxy, for example), or irradiation such as electron beam
processing.

Thermoset materials are usually liquid or malleable prior to curing and designed to be
molded into their final form, or used as adhesives. Others are solids like that of the
molding compound used in semiconductors and integrated circuits (IC's).

Process

The curing process transforms the resin into a plastic or rubber by a cross-linking
process. Energy and/or catalysts are added that cause the molecular chains to react at
chemically active sites (unsaturated or epoxy sites, for example), linking into a rigid, 3-D
structure. The cross-linking process forms a molecule with a larger molecular weight,
resulting in a material with a higher melting point. During the reaction, the molecular
weight has increased to a point so that the melting point is higher than the surrounding
ambient temperature, the material forms into a solid material.
Uncontrolled reheating of the material results in reaching the decomposition temperature
before the melting point is obtained. Therefore, a thermoset material cannot be melted
and re-shaped after it is cured. This implies that thermosets cannot be recycled, except as
filler material.[1]

Statistics

Thermoset materials are generally stronger than thermoplastic materials due to this 3-D
network of bonds, and are also better suited to high-temperature applications up to the
decomposition temperature.

Examples

Some examples of thermosets are:

• Polyester fiberglass systems: (SMC Sheet molding compounds and BMC Bulk
molding compounds)
• Vulcanized rubber
• Bakelite, a phenol-formaldehyde resin (used in electrical insulators and
plasticware)
• Duroplast, similar to Bakelite
• Urea-formaldehyde foam (used in plywood, particleboard and medium-density
fibreboard)
• Melamine resin (used on worktop surfaces)
• Epoxy resin (used as an adhesive and in fibre reinforced plastics such as glass
reinforced plastic and graphite-reinforced plastic)
• Polyimides (used in printed circuit boards and in body parts of modern airplanes)
• Mold or Mold Runners (the black plastic part in Integrated Circuits (IC) or
semiconductors)

Some methods of molding thermosets are:

• Reactive injection molding (used for objects like milk bottle crates)
• Extrusion molding (used for making pipes, threads of fabric and insulation for
electrical cables)
• Compression molding (used to shape most thermosetting plastics)
• Spin casting (used for producing fishing lures and jigs, gaming miniatures,
figurines, emblems as well as production and replacement parts
Thermoplastics vs Thermosetting
Thermoplastics and thermosetting plastics are terms that describe how a polymer reacts to
heat. All plastics, whether made by addition or condensation polymerization, can be
divided into two groups: thermoplastics and thermosetting plastics. Thermoplastics can
be repeatedly softened by heating and hardened by cooling. Thermosetting plastics, on
the other hand, harden permanently after being heated once.

 The Difference - Weak Van Der Waal Forces


The reason for the difference in response to heat between thermoplastics and
thermosetting plastics lies in the chemical structures of the plastics. Thermoplastic
molecules, which are linear or slightly branched, do not chemically bond with each
other when heated. Instead, thermoplastic chains are held together by weak van der
Waal forces (weak attractions between the molecules) that cause the long molecular
chains to clump together like piles of entangled spaghetti. Thermoplastics can be
heated and cooled, and consequently softened and hardened, repeatedly, like candle
wax. For this reason, thermoplastics can be remolded and reused almost indefinitely.

 Thermosetting Plastics
Thermosetting plastics consist of chain molecules that chemically bond, or cross-link,
with each other when heated. When thermosetting plastics cross-link, the molecules
create a permanent, three-dimensional network that can be considered one giant
molecule. Once cured, thermosetting plastics cannot be remelted, in the same way
that cured concrete cannot be reset. Consequently, thermosetting plastics are often
used to make heat-resistant products, because these plastics can be heated to
temperatures of 260° C (500° F) without melting.

 Thermoplastics
The different molecular structures of thermoplastics and thermosetting plastics allow
manufacturers to customize the properties of commercial plastics for specific
applications. Because thermoplastic materials consist of individual molecules,
properties of thermoplastics are largely influenced by molecular weight. For instance,
increasing the molecular weight of a thermoplastic material increases its tensile
strength, impact strength, and fatigue strength (ability of a material to withstand
constant stress). Conversely, because thermosetting plastics consist of a single
molecular network, molecular weight does not significantly influence the properties
of these plastics. Instead, many properties of thermosetting plastics are determined by
adding different types and amounts of fillers and reinforcements, such as glass fibers.
 The Processes of Making Plastics
The process of forming plastic resins into plastic products is the basis of the plastics
industry. Many different processes are used to make plastic products, and in each
process, the plastic resin must be softened or sufficiently liquefied to be shaped.
Although some processes are used to manufacture both thermoplastics and
thermosetting plastics, certain processes are specific to forming thermoplastics.

 Injection Molding
Injection molding uses a piston or screw to force plastic resin through a heated tube
into a mold, where the plastic cools and hardens to the shape of the mold. The mold is
then opened and the plastic cast removed. Thermoplastic items made by injection
molding include toys, combs, car grills, and various
containers.

 Extrusion
Extrusion is a continuous process, as opposed to all other plastic production
processes, which start over at the beginning of the process after each new part is
removed from the mold. In the extrusion process, plastic pellets are first heated in a
long barrel. In a manner similar to that of a pasta-making or sausage-stuffing
machine, a rotating screw then forces the heated plastic through a die (device used for
forming material) opening of the desired shape.

As the continuous plastic form emerges from the die opening, it is cooled and
solidified, and the continuous plastic form is then cut to the desired length. Plastic
products made by extrusion include garden hoses, drinking straws, pipes, and ropes.
Melted thermoplastic forced through extremely fine die holes can be cooled and
woven into fabrics for clothes, curtains, and carpets.

 Blow Molding
Blow molding is used to form bottles and other containers from soft, hollow
thermoplastic tubes. First a mold is fitted around the outside of the softened
thermoplastic tube, and then the tube is heated. Next, air is blown into the softened
tube (similar to inflating a balloon), which forces the outside of the softened tube to
conform to the inside walls of the mold. Once the plastic cools, the mold is opened
and the newly molded container is removed. Blow molding is used to make many
plastic containers, including soft-drink bottles, jars, detergent bottles, and storage
drums.
Natural rubber
Natural rubber is an elastomer (an elastic hydrocarbon polymer) that was originally
derived from a milky colloidal suspension, or latex, found in the sap of some plants. The
purified form of natural rubber is the chemical polyisoprene which can also be produced
synthetically. Natural rubber is used extensively in many applications and products as is
synthetic rubber. The entropy model of rubber was developed in 1934 by Werner Kuhn.

Latex being collected from a tapped rubber tree

Varieties

The major commercial source of natural rubber latex is the Para rubber tree (Hevea
brasiliensis), a member of the spurge family, Euphorbiaceae. This is largely because it
responds to wounding by producing more latex.

Other plants containing latex include Gutta-Percha (Palaquium gutta),[1] rubber fig (Ficus
elastica), Panama rubber tree (Castilla elastica), spurges (Euphorbia spp.), lettuce,
common dandelion (Taraxacum officinale), Russian dandelion (Taraxacum kok-saghyz),
Scorzonera tau-saghyz, and Guayule (Parthenium argentatum). Although these have not
been major sources of rubber, Germany attempted to use some of these during World
War II when it was cut off from rubber supplies[citation needed]. These attempts were later
supplanted by the development of synthetic rubbers. To distinguish the tree-obtained
version of natural rubber from the synthetic version, the term gum rubber is sometimes
used.
Discovery of commercial potential

Charles Marie de La Condamine is credited with introducing samples of rubber to the


Académie Royale des Sciences of France in 1736.[2] In 1751, he presented a paper by
François Fresneau to the Académie (eventually published in 1755) which described many
of the properties of rubber. This has been referred to as the first scientific paper on
rubber.[2]

The para rubber tree initially grew in South America, and the first European to return to
Portugal from Brazil with samples of water-repellent rubberized cloth so shocked people
that he was brought to court on the charge of witchcraft. When samples of rubber first
arrived in England, it was observed by Joseph Priestley, in 1770, that a piece of the
material was extremely good for rubbing out pencil marks on paper, hence the name
rubber.

South America remained the main source of what limited amount of latex rubber was
consumed during much of the 19th century. However in 1876, Henry Wickham gathered
thousands of seeds from Brazil, and these were germinated in Kew Gardens, UK. The
seedlings were then sent to Ceylon (Sri Lanka), Indonesia, Singapore and British Malaya.
Malaya (now Malaysia) was later to become the biggest producer of rubber. About 100
years ago, the Congo Free State in Africa was also a significant source of natural rubber
latex, mostly gathered by forced labor. Liberia and Nigeria also started production of
rubber.

In India, commercial cultivation of natural rubber was introduced by the British Planters,
although the experimental efforts to grow rubber on a commercial scale in India were
initiated as early as 1873 at the Botanical Gardens, Kolkata. The first commercial Hevea
plantations in India were established at Thattekadu in Kerala in 1902.

Properties
Rubber latex.

Rubber exhibits unique physical and chemical properties. Rubber's stress-strain behavior
exhibits the Mullins effect, the Payne effect and is often modeled as hyperelastic. Rubber
strain crystallizes.

Owing to the presence of a double bond in each and every repeat unit, natural rubber is
sensitive to ozone cracking.
Solvents

There are two main solvents for rubber: turpentine and naphtha (petroleum). The former
has been in use since 1763 when Francois Fresnau made the discovery. Giovanni
Fabronni is credited with the discovery of naphtha as a rubber solvent in 1779. Because
rubber does not dissolve easily, the material is finely divided by shredding prior to its
immersion.An ammonia solution can be used to prevent the coagulation of raw latex
while it is being transported from its collection site.

Chemical makeup

Natural rubber is a polymer of isoprene - most often cis-1,4-polyisoprene - with a


molecular weight of 100,000 to 1,000,000. Typically, a few percent of other materials,
such as proteins, fatty acids, resins and inorganic materials are found in natural rubber.
Polyisoprene is also created synthetically, producing what is sometimes referred to as
"synthetic natural rubber".

Some natural rubber sources called gutta percha are composed of trans-1,4-polyisoprene,
a structural isomer which has similar, but not identical properties.

Natural rubber is an elastomer and a thermoplastic. However, it should be noted that as


the rubber is vulcanized it will turn into a thermoset. Most rubber in everyday use is
vulcanized to a point where it shares properties of both; i.e., if it is heated and cooled, it is
degraded but not destroyed.

Elasticity

In most elastic materials, such as metals used in springs, the elastic behavior is caused by
bond distortions. When force is applied, bond lengths deviate from the (minimum energy)
equilibrium and strain energy is stored electrostatically. Rubber is often assumed to
behave in the same way, but it turns out this is a poor description. Rubber is a curious
material because, unlike metals, strain energy is stored thermally.

In its relaxed state rubber consists of long, coiled-up polymer chains that are interlinked
at a few points. Between a pair of links each monomer can rotate freely about its
neighbour. This gives each section of chain leeway to assume a large number of
geometries, like a very loose rope attached to a pair of fixed points. At room temperature
rubber stores enough kinetic energy so that each section of chain oscillates chaotically,
like the above piece of rope being shaken violently.

When rubber is stretched the "loose pieces of rope" are taut and thus no longer able to
oscillate. Their kinetic energy is given off as excess heat. Therefore, the entropy
decreases when going from the relaxed to the stretched state, and it increases during
relaxation. This change in entropy can also be explained by the fact that a tight section of
chain can fold in fewer ways (W) than a loose section of chain, at a given temperature
(nb. entropy is defined as S=k*ln(W)). Relaxation of a stretched rubber band is thus
driven by an increase in entropy, and the force experienced is not electrostatic, rather it is
a result of the thermal energy of the material being converted to kinetic energy. Rubber
relaxation is endothermic, and for this reason the force exerted by a stretched piece of
rubber increases with temperature (metals, for example, become softer as temperature
increases). The material undergoes adiabatic cooling during contraction. This property of
rubber can easily be verified by holding a stretched rubber band to your lips and relaxing
it.

Stretching of a rubber band is in some ways equivalent to the compression of an ideal


gas, and relaxation is equivalent to its expansion. Note that a compressed gas also
exhibits "elastic" properties, for instance inside an inflated car tire. The fact that
stretching is equivalent to compression may seem somewhat counter-intuitive, but it
makes sense if rubber is viewed as a one-dimensional gas. Stretching reduces the "space"
available to each section of chain.

Vulcanization of rubber creates more disulfide bonds between chains so it makes each
free section of chain shorter. The result is that the chains tighten more quickly for a given
length of strain. This increases the elastic force constant and makes rubber harder and
less extendable.

When cooled below the glass transition temperature, the quasi-fluid chain segments
"freeze" into fixed geometries and the rubber abruptly loses its elastic properties, though
the process is reversible. This is a property it shares with most elastomers. At very cold
temperatures rubber is actually rather brittle; it will break into shards when struck or
stretched. This critical temperature is the reason that winter tires use a softer version of
rubber than normal tires. The failing rubber o-ring seals that contributed to the cause of
the Challenger disaster were thought to have cooled below their critical temperature. The
disaster happened on an unusually cold day.

Current sources

Close to 21 million tons of rubber were produced in 2005 of which around 42% was
natural. Since the bulk of the rubber produced is the synthetic variety which is derived
from petroleum, the price of even natural rubber is determined to a very large extent by
the prevailing global price of crude oil[citation needed]. Today Asia is the main source of
natural rubber, accounting for around 94% of output in 2005. The three largest producing
countries (Indonesia, Malaysia and Thailand) together account for around 72% of all
natural rubber production.
Cultivation

Rubber is generally cultivated in large plantations. See the coconut shell used in
collecting latex, in plantations in Kerala, India

Rubber latex is extracted from Rubber trees. The economic life period of rubber trees in
plantations is around 32 years – up to 7 years of immature phase and about 25 years of
productive phase.

The soil requirement of the plant is generally well-drained weathered soil consisting of
laterite, lateritic types, sedimentary types, nonlateritic red or alluvial soils.

The climatic conditions for optimum growth of Rubber trees consist of (a) Rainfall of
around 250 cm evenly distributed without any marked dry season and with at least 100
rainy days per annum (b) Temperature range of about 20°C to 34°C with a monthly mean
of 25°C to 28°C (c) High atmospheric humidity of around 80% (d) Bright sunshine
amounting to about 2000 hours per annum at the rate of 6 hours per day throughout the
year and (e) Absence of strong winds.

Many high-yielding clones have been developed for commercial planting. These clones
yield more than 2,000 kilograms of dry Rubber per hectare per annum, when grown
under ideal conditions.

Collection

In places like Kerala, where coconuts are in abundance, the shell of half a coconut is used
as the collection container for the latex but glazed pottery or aluminium cups are more
common elsewhere. The cups are supported by a wire that encircles the tree.This wire
incorporates a spring so that it can stretch as the tree grows. The latex is led into the cup
by a galvanised "spout" that has been knocked into the bark. Tapping normally takes
place early in the morning when the internal pressure of the tree is highest. A good tapper
can tap a tree every 20 seconds on a standard half-spiral system and a common daily
"task" size is between 450 and 650 trees. Trees are usually tapped alternate or third daily
although there are many variations in timing, length and number of cuts. The latex, which
contains 25 - 40% dry rubber, is in the bark so the tapper must avoid cutting right through
to the wood or the growing cambial layer will be damaged and the renewing bark will be
badly deformed making later tapping difficult. It is usual to tap a pannel at least twice,
sometimes three times, during the trees' life. The economic life of the tree depends on
how well the tapping is carried out as the critical factor is bark consumption. A standard
in Malaysia for alternate daily tapping is 25 cm (vertical) bark consumption per annum.
The latex tubes in the bark ascend in a spiral to the right. For this reason, tapping cuts
usually ascend to the left to cut more tubes.
A tree woman in Sri Lanka in the process of harvesting rubber

The trees will drip latex for about four hours, stopping as latex coagulates naturally on
the tapping cut thus blocking the latex tubes in the bark. Tappers usually rest and have a
meal after finishing their tapping work then start collecting the latex at about midday.
Some trees will continue to drip after the collection and this leads to a small amount of
cup lump which is collected at the next tapping. The latex that coagulates on the cut is
also collected as tree lace. Tree lace and cup lump together account for 10 - 20% of the
dry rubber produced.

The latex can be collected in its liquid state. It is sometimes necessary to add a few drops
of ammonia solution to the cup, or to the transport tank, to prevent precoagulation of the
latex before it reaches the factory. It can also be left in the cup to coagulate naturally into
cup lump for collection before the next tapping, although this will produce a lower grade
of product.

Latex is generally processed into either latex concentrate for manufacture of dipped
goods or it can be coagulated under controlled, clean conditions using formic acid. The
coagulated latex can then be processed into the higher grade technically specified block
rubbers such as TSR3L or TSRCV or used to produce Ribbed Smoke Sheet grades.

Naturally coagulated rubber (cup lump) is used in the manufacture of TSR10 and TSR20
grade rubbers. The processing of the rubber for these grades is basically a size reduction
and cleaning process in order to remove contamination and prepare the material for the
final stage drying.

The dried material is then baled and palletized for shipment.


Uses

Compression molded (cured) rubber boots before the flashes


are removed.

The use of rubber is widespread, ranging from household to


industrial products, entering the production stream at the
intermediate stage or as final products. Tires and tubes are
the largest consumers of rubber, accounting for around 56%
total consumption in 2005. The remaining 44% are taken up
by the general rubber goods (GRG) sector, which includes all
products except tires and tubes.

Pre-historical uses

The first use of rubber was natural latex from the Hevea Tree in 1600 BC by the Ancient
Mayans[citation needed]. They boiled the harvested latex to make a ball for sport

Manufacturing

Other significant uses of rubber are door and window profiles, hoses, belts, matting,
flooring and dampeners (anti-vibration mounts) for the automotive industry in what is
known as the "under the bonnet" products. Gloves (medical, household and industrial) are
also large consumers of toy balloons and rubber, although the type of rubber used is that
of the concentrated latex. Significant tonnage of rubber is used as adhesives in many
manufacturing industries and products, although the two most noticeable are the paper
and the carpet industry. Rubber is also commonly used to make rubber bands and pencil
erasers.

Textile applications

Additionally, rubber produced as a fiber sometimes called elastic, has significant value
for use in the textile industry because of its excellent elongation and recovery properties.
For these purposes, manufactured rubber fiber is made as either an extruded round fiber
or rectangular fibers that are cut into strips from extruded film. Because of its low dye
acceptance, feel and appearance, the rubber fiber is either covered by yarn of another
fiber or directly woven with other yarns into the fabric. In the early 1900’s, for example,
rubber yarns were used in foundation garments. While rubber is still used in textile
manufacturing, its low tenacity limits its use in lightweight garments because latex lacks
resistance to oxidizing agents and is damaged by aging, sunlight, oil, and perspiration.
Seeking a way to address these shortcomings, the textile industry has turned to Neoprene
(polymer form of Chloroprene), a type of synthetic rubber as well as another more
commonly used elastomer fiber, spandex (also known as elastane), because of their
superiority to rubber in both strength and durability.

Vulcanization
Main article: Vulcanization

Natural rubber is often vulcanized, a process by which the rubber is heated and sulfur,
peroxide or bisphenol are added to improve resilience and elasticity, and to prevent it
from perishing. Vulcanization greatly improved the durability and utility of rubber from
the 1830s on[citation needed]. The development of vulcanization is most closely associated with
Charles Goodyear[3]. Carbon black is often used as an additive to rubber to improve its
strength, especially in vehicle tires.

Allergic reactions
Main article: Latex allergy

Some people have a serious latex allergy, and exposure to certain natural rubber latex
products such as latex gloves can cause anaphylactic shock. Guayule latex is
hypoallergenic and is being researched as a substitute to the allergy-inducing Hevea
latexes. Unlike the sappable Hevea tree, these relatively small shrubs must be harvested
whole and latex extracted from each cell.

Some allergic reactions are not from the latex but from residues of other ingredients used
to process the latex into clothing, gloves, foam, etc. These allergies are usually referred to
as multiple chemical sensitivity (MCS).

Objective Questions

1. A copolymer has the following repeat unit:


-{CH2-CHCI- CH2-CH=CH- CH2}-
Which pair of monomers could be used to make this polymer?

A. CH2 = CHCI and CH2 = CH2


B. CH2 = CHCI and CH2 = CH-CH= CH2
C. CH3 - CH2CI and CH3 – CH= CH- CH3
D. CH2 =CCI-CH= CH2 and CH2 = CH2

Answer: B

2 Natural rubber is made up of isoprene monomer units with the structural formula:
CH2 = CH(CH3) CH= CH2
Which of the following about rubber and its monomer is not true?

A The IUPAC name for isoprene is 2- methylbuta- 1,3- diene

B Raw rubber is soft and sticky when warmed

C Bulk rubber is a mixture of cis- and trans- polyisoprene

D Bulk rubber can be strenghten by vulcanization

Answer: C
(Bulk rubber is raw natural rubber.Natural rubber consists only of cis- polyisoprene
molecular chains)

3. The sequences of reactions below shows how benzene is used to manufacture poly
(phenylethene)

benzene phenylethane phenylethene poly(phenylethene)


Which of the following statements can be used to explain the above process?

1. Using a suitable catalyst, ethene can be substituted into benzene to form


phenylbenzene

2. Phenylethene is obtained from phenylethene by dehydrogenatio

3. The repeat unit in poly(phenylethene) is

Answer: B [ only1 and 2 are correct]


( Another name for poly(phenylethene) is polystrene)

4. Clerfilm is manufactured from a polymer made by copolymerizing CH2 = CHCI with.


CH2 =CCI2 in a regular ‘ head to tall’ linkage, where CH2 is taken as the ‘ head’ of the
monomer. Which of the following could represent parts of the polymer chain in
‘clearfilm’?

1. -CHCI-CH2-CCI2-

2. - CCI2- CCI2- CH2- CHCI

3. - CH2- CHCI- CCI2

Answer: A [ only 1 is correct]


(Only one arrangement for the length of the polymer is possible, - CH2- CHCI-
CH2- CCI2- CH2- CHCI)

5. A part of a certain polymeric chain is shown below.


Which of the following statements about this polyner are correct?

1. Its monomer is

2. It’s empirical formula is C2H2CI

3. It has geometrical isomers

Answer: Conly 2 and 3 are correct]

(Its monomer is )

6. PET or polyethreneterephthalate is the most widely used packing polymer. PET can
be categorized as a

1. polyethene type polymer

2. condensation polymer

3. polyster

Answer: D [1,2 and 3 are correct]

Subjective Questions

1. The following scheme shows how nylon 6,6 is manufactured starting from phenol
H2
OH OH O
Catalyst J

K L

Oxidation

CO2H

(CH2)4 + M Nylon 6,6

CO2H

a.) Write a balanced equation for the reaction between phenol and hydrogen.

OH OH

+3H2

b.) Name or write the formula for the


i.) Catalyst J
Answer: Nickel / Platinum

ii.) Compond M
Answer: Hexane- 1,6 – diamine

c.) Draw the structural formula of the repeat unit in nylon 6,6

d.) Describe a simple chemical test to distingiuish between the compounds K and L
Answer: Add phosphorus pentachloride. K produces dense white fumes.

OH CI

+ 5PCI5 + POCI3 + HCI White fumes


No changes are seen with L
Comment:
Tests on the carbohynl group in L can also be accepted

e.) Which of the 2 componds K and L will have a higher boiling point? Explain your
answer?
Answer: K. The alcohol K is capable of self hydrogen bonding which raises its
boiling point. L is not capable of hydrogen bonding.

2.) a) Polymers can be divided into thermosets, thermoplastics and elastomers.


i) What is a thermoplastic polymer?
Answer: Thermoplastic polymers soften on warning and can be remoulded.
Structurally, thermoplastic polymers do not have cross-link between
polymeric chains.

ii.) Give an example of a thermoplastic polymer?


Answer: PE / PS / PVS / PET / NYLON

b.) Natural rubber is a elastomer.


i.) What is meant by the term elastomer?
Answer: An elastomer is a polymer that can be streched to at least twice its
original length and rapidly contracts to its original length when released .

ii.) What is the monomer unit in natural rubber?


Answer: The monomer unit is natural rubber is 2- methylbutadiene or more
commonly known as isoprene.

c.) Poly( buta-1,3 – diene) is a synthetic rubber. It exsists as a mixture of


geometrical isomers, is sticky and is not useful material. What is added to
synthetic rubber to give it more resilient properties?
Answer: Sulphur

d.) SBR is a synthetic rubber used in the automobile industry. Name the monomers
that are used to make SBR.
Answer: Phenylethene / Styrene and buta – 1,3 – diene

3.) a.) Explain the following terms, using suitable examples .


i.) Addition Polymerization
Answer: Addition Polymerization is the reaction of alkenes or functionally –
substituted alkene monomers by addition to form polymers. It proceeds by
way of chain-growth mechanisms involving reactive intermediates which my
be a free radical, an anion or a cation.
ii.) Condensation Polymerization
Answer: Condensation Polymerization is the formation of a polymer by
functional group reaction which often occur with loss of a small by- product
such as water. Reaction between the functional groups usually gives it an
alternate structure.

b.) How do the properties of a polymer change under the follwing conditions?
i.) Lenghtening the polymer
Answer: The softening temperature of the polymer can be raised by
lengthening. Soft and sticky polymers can be made more rigid.

ii.) Forming cross-links


Answer: Moderate cross-linking results in a more elastic and resilient polymer.
Generally, a cross-linked polymer is thermoset and cannot be reshapad

iii.) Adding a polymer


Answer: Adding a copolymer blends the properties of polymers which are
often better than either of the orginal polymers.

iv.) Adding a plasticizer


Answer: A plastizer is added to incrase the flexbility and durability of a
polymer.

CONCLUSION

We have attempted to provide an overview of the medical device uses of biodegradable


polymers. While sutures were the first commercial product and still account for the vast
majority of all sales, a variety of products are now on the market for an expanding range
of applications, with others certain to appear in the next decade.
What is it about these materials that makes them so attractive to the device industry?
First, in this conservative field, where devices serve critical, perhaps life-and-death,
functions, the industry is slow to accept new materials or new designs. The polymers
prepared from these materials, particularly lactide and glycolide, have a long history of
safe and effective use. Building on this solid foundation, researchers will continue to
evaluate these materials, taking advantage of the wide range of properties that can be
obtained in polymers built with relatively few monomer units. We expect that, in the
future even more than today, device designers and physicians will have available a wealth
of products using biodegradable polymers that will help speed patient recovery and
eliminate follow-up surgeries.
"Polymers." University of Illinois, Urbana/Champaign,Materials Science and
Technology Department (Web site).
<http://matse1.mse.uiuc.edu/~tw/polymers/polymers.html> (June 5, 2001)

"Polymers and Liquid Crystals." Case Western ReserveUniversity (Web site).


<http://abalone.cwru.edu/> (June 5, 2001)

Plastics.com (Web site). <http://www.plastics.com> (June5, 2001)

Plastics 101." Plastics Resource (Web site).


<http://www2.plasticsresource.com/plastics_101/> (June 5, 2001

http://www.smithsonianmag.com/science-nature/plastic.html

^ http://www.manufacturingtalk.com/news/trj/trj110.html

Potrebbero piacerti anche