You are on page 1of 17

Polymer

From Wikipedia, the free encyclopedia

Jump to navigationJump to search

For other uses, see Polymer (disambiguation).

Appearance of real linear polymer chains as recorded using an atomic force microscope on a surface,
under liquid medium. Chain contour length for this polymer is ~204 nm; thickness is ~0.4 nm.[1]

IUPAC definition

A polymer is a substance composed of macromolecules.[2] A macromolecule is a molecule of high


relative molecular mass, the structure of which essentially comprises the multiple repetition of units
derived, actually or conceptually, from molecules of low relative molecular mass.[3]

Condensed matter physics

QuantumPhaseTransition.svg

Phases · Phase transition · QCP

States of matter

Phase phenomena

Electronic phases

Electronic phenomena

Magnetic phases

Quasiparticles

Soft matter

Scientists

vte

A polymer (/ˈpɒlɪmər/;[4][5] Greek poly-, "many" + -mer, "part") is a substance or material consisting
of very large molecules, or macromolecules, composed of many repeating subunits.[6] Due to their
broad spectrum of properties,[7] both synthetic and natural polymers play essential and ubiquitous
roles in everyday life.[8] Polymers range from familiar synthetic plastics such as polystyrene to
natural biopolymers such as DNA and proteins that are fundamental to biological structure and
function. Polymers, both natural and synthetic, are created via polymerization of many small
molecules, known as monomers. Their consequently large molecular mass, relative to small
molecule compounds, produces unique physical properties including toughness, high elasticity,
viscoelasticity, and a tendency to form amorphous and semicrystalline structures rather than
crystals.

The term "polymer" derives from the Greek word πολύς (polus, meaning "many, much") and μέρος
(meros, meaning "part"), and refers to large molecules whose structure is composed of multiple
repeating units, from which originates a characteristic of high relative molecular mass and attendant
properties.[3] The units composing polymers derive, actually or conceptually, from molecules of low
relative molecular mass.[3] The term was coined in 1833 by Jöns Jacob Berzelius, though with a
definition distinct from the modern IUPAC definition.[9][10] The modern concept of polymers as
covalently bonded macromolecular structures was proposed in 1920 by Hermann Staudinger,[11]
who spent the next decade finding experimental evidence for this hypothesis.[12]

Polymers are studied in the fields of polymer science (which includes polymer chemistry and
polymer physics), biophysics and materials science and engineering. Historically, products arising
from the linkage of repeating units by covalent chemical bonds have been the primary focus of
polymer science. An emerging important area now focuses on supramolecular polymers formed by
non-covalent links. Polyisoprene of latex rubber is an example of a natural polymer, and the
polystyrene of styrofoam is an example of a synthetic polymer. In biological contexts, essentially all
biological macromolecules—i.e., proteins (polyamides), nucleic acids (polynucleotides), and
polysaccharides—are purely polymeric, or are composed in large part of polymeric components.

Cartoon schematic of polymer molecules

Contents

1 Common examples

1.1 Natural

1.2 Synthetic

2 History

3 Synthesis

3.1 Biological synthesis

3.2 Modification of natural polymers

4 Structure

4.1 Monomers and repeat units

4.2 Microstructure

4.2.1 Polymer architecture

4.2.2 Chain length

4.2.3 Monomer arrangement in copolymers

4.2.4 Tacticity

4.3 Morphology
4.3.1 Crystallinity

4.3.2 Chain conformation

5 Properties

5.1 Mechanical properties

5.1.1 Tensile strength

5.1.2 Young's modulus of elasticity

5.2 Transport properties

5.3 Phase behavior

5.3.1 Crystallization and melting

5.3.2 Glass transition

5.3.3 Mixing behavior

5.3.4 Inclusion of plasticizers

5.4 Chemical properties

5.5 Optical properties

5.6 Electrical properties

6 Applications

7 Standardized nomenclature

8 Characterization

9 Degradation

9.1 Product failure

10 See also

11 References

12 Bibliography

13 External links

Common examples

See also: Polymer classes

Structure of a styrene-butadiene chain, from a molecular simulation.

Polymers are of two types: naturally occurring and synthetic or man made.

Natural
Natural polymeric materials such as hemp, shellac, amber, wool, silk, and natural rubber have been
used for centuries. A variety of other natural polymers exist, such as cellulose, which is the main
constituent of wood and paper.

Synthetic

The list of synthetic polymers, roughly in order of worldwide demand, includes polyethylene,
polypropylene, polystyrene, polyvinyl chloride, synthetic rubber, phenol formaldehyde resin (or
Bakelite), neoprene, nylon, polyacrylonitrile, PVB, silicone, and many more. More than 330 million
tons of these polymers are made every year (2015).[13]

Most commonly, the continuously linked backbone of a polymer used for the preparation of plastics
consists mainly of carbon atoms. A simple example is polyethylene ('polythene' in British English),
whose repeating unit is based on ethylene monomer. Many other structures do exist; for example,
elements such as silicon form familiar materials such as silicones, examples being Silly Putty and
waterproof plumbing sealant. Oxygen is also commonly present in polymer backbones, such as those
of polyethylene glycol, polysaccharides (in glycosidic bonds), and DNA (in phosphodiester bonds).

History

See also: Polymer science

Polymers have been essential components of commodities since the early days of humankind. The
use of wool (keratin), cotton and linen fibres (cellulose) for garments, paper reed (cellulose) for
paper are just a few examples of how our ancestors exploited polymer-containing raw materials to
obtain artefacts. The latex sap of “caoutchouc” trees (natural rubber) reached Europe in the 16th
century from South America long after the Olmec, Maya and Aztec had started using it as a material
to make balls, waterproof textiles and containers.[14]

The chemical manipulation of polymers dates back to the 19th century, although at the time the
nature of these species was not understood. The behaviour of polymers was initially rationalised
according to the theory proposed by Thomas Graham which considered them as colloidal aggregates
of small molecules held together by unknown forces.

Notwithstanding the lack of theoretical knowledge, the potential of polymers to provide innovative,
accessible and cheap materials was immediately grasped. The work carried out by Braconnot,
Parkes, Ludersdorf, Hayard and many others on the modification of natural polymers determined
many significant advances in the field.[15] Their contributions led to the discovery of materials such
as celluloid, galalith, parkesine, rayon, vulcanised rubber and, later, Bakelite: all materials that
quickly entered industrial manufacturing processes and reached households as garments
components (e.g., fabrics, buttons), crockery and decorative items.
In 1920, Hermann Staudinger published his seminal work “Über Polymerisation”,[16] in which he
proposed that polymers were in fact long chains of atoms linked by covalent bonds. His work was
debated at length, but eventually it was accepted by the scientific community. Because of this work,
Staudinger was awarded the Nobel Prize in 1953.[17]

After the 1930s polymers entered a golden age during which new types were discovered and quickly
given commercial applications, replacing naturally-sourced materials. This development was fuelled
by an industrial sector with a strong economical drive and it was supported by a wide academic
community that contributed with innovative synthesis of monomers from cheaper raw materials,
more efficient polymerisation processes, improved techniques for polymer characterisation and
advanced theoretical understanding of polymers.[15]

Since 1953, six Nobel prizes were awarded in the area of polymer science, excluding those for
research on biological macromolecules. This further testifies its impact on modern science and
technology. As Lord Todd summarised it in 1980, “I am inclined to think that the development of
polymerization is perhaps the biggest thing that chemistry has done, where it has had the biggest
effect on everyday life”.[19]

Synthesis

Main article: Polymerization

A classification of the polymerization reactions

Polymerization is the process of combining many small molecules known as monomers into a
covalently bonded chain or network. During the polymerization process, some chemical groups may
be lost from each monomer. This happens in the polymerization of PET polyester. The monomers are
terephthalic acid (HOOC—C6H4—COOH) and ethylene glycol (HO—CH2—CH2—OH) but the
repeating unit is —OC—C6H4—COO—CH2—CH2—O—, which corresponds to the combination of
the two monomers with the loss of two water molecules. The distinct piece of each monomer that is
incorporated into the polymer is known as a repeat unit or monomer residue.

Synthetic methods are generally divided into two categories, step-growth polymerization and chain
polymerization.[20] The essential difference between the two is that in chain polymerization,
monomers are added to the chain one at a time only,[21] such as in polystyrene, whereas in step-
growth polymerization chains of monomers may combine with one another directly,[22] such as in
polyester. Step-growth polymerization can be divided into polycondensation, in which low-molar-
mass by-product is formed in every reaction step, and polyaddition.
Example of chain polymerization: Radical polymerization of styrene, R. is initiating radical, P. is
another polymer chain radical terminating the formed chain by radical recombination

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. Laboratory synthesis of
biopolymers, especially of proteins, is an area of intensive research.

Biological synthesis

Main article: Biopolymer

Microstructure of part of a DNA double helix biopolymer

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. There are
other biopolymers such as rubber, suberin, melanin, and lignin.

Modification of natural polymers

Naturally occurring polymers such as cotton, starch, and rubber were familiar materials for years
before synthetic polymers such as polyethene and perspex appeared on the market. 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 sulfur. Ways in
which polymers can be modified include oxidation, cross-linking, and endcapping.

Structure

The structure of a polymeric material can be described at different length scales, from the sub-nm
length scale up to the macroscopic one. There is in fact a hierarchy of structures, in which each stage
provides the foundations for the next one.[23] The starting point for the description of the structure
of a polymer is the identity of its constituent monomers. Next, the microstructure essentially
describes the arrangement of these monomers within the polymer at the scale of a single chain. The
microstructure determines the possibility for the polymer to form phases with different
arrangements, for example through crystallization, the glass transition or microphase
separation.[24] These features play a major role in determining the physical and chemical properties
of a polymer.

Monomers and repeat units


The identity of the repeat units (monomer residues, also known as "mers") comprising a polymer is
its first and most important attribute. Polymer nomenclature is generally based upon the type of
monomer residues comprising the polymer. A polymer which contains only a single type of repeat
unit is known as a homopolymer, while a polymer containing two or more types of repeat units is
known as a copolymer.[25] A terpolymer is a copolymer which contains three types of repeat
units.[26]

Polystyrene is composed only of styrene-based repeat units, and is classified as a homopolymer.


Polyethylene terephthalate, even though produced from two different monomers (ethylene glycol
and terephthalic acid), is usually regarded as a homopolymer because only one type of repeat unit is
formed. Ethylene-vinyl acetate contains more than one variety of repeat unit and is a copolymer.
Some biological polymers are composed of a variety of different but structurally related monomer
residues; for example, polynucleotides such as DNA are composed of four types of nucleotide
subunits.

Homopolymers and copolymers (examples)

Polystyrene skeletal.svg

Poly(dimethylsiloxan).svg

PET.svg

Styrol-Butadien-Kautschuk.svg

Homopolymer polystyrene Homopolymer polydimethylsiloxane, a silicone. The main chain is


formed of silicon and oxygen atoms. The homopolymer polyethylene terephthalate has only one
repeat unit. Copolymer styrene-butadiene rubber: The repeat units based on styrene and 1,3-
butadiene form two repeating units, which can alternate in any order in the macromolecule, making
the polymer thus a random copolymer.

A polymer molecule containing ionizable subunits is known as a polyelectrolyte or ionomer.

Microstructure

Main article: Microstructure

The microstructure of a polymer (sometimes called configuration) relates to the physical


arrangement of monomer residues along the backbone of the chain.[27] These are the elements of
polymer structure that require the breaking of a covalent bond in order to change. Various polymer
structures can be produced depending on the monomers and reaction conditions: A polymer may
consist of linear macromolecules containing each only one unbranched chain. In the case of
unbranched polyethylene, this chain is a long-chain n-alkane. Linear polymers may fold into diverse
conformations with distinct circuit topology. There are also branched macromolecules with a main
chain and side chains, in the case of polyethylene the side chains would be alkyl groups. In particular
unbranched macromolecules can be in the solid state semi-crystalline, crystalline chain sections
highlighted red in the figure below.
While branched and unbranched polymers are usually thermoplastics, many elastomers have a
wide-meshed cross-linking between the "main chains". Close-meshed crosslinking, on the other
hand, leads to thermosets. Cross-links and branches are shown as red dots in the figures. Highly
branched polymers are amorphous and the molecules in the solid interact randomly.

Polymerstruktur-linear.svg

linear, unbranched macromolecule Polymerstruktur-verzweigt.svg

branched macromolecule Polymerstruktur-teilkristallin.svg

semi-crystalline structure of an unbranched polymer Polymerstruktur-weitmaschig vernetzt.svg

slightly cross-linked polymer (elastomer) Polymerstruktur-engmaschig vernetzt.svg

highly cross-linked polymer (thermoset)

Polymer architecture

Main article: Polymer architecture

Branch point in a polymer

An important microstructural feature of a polymer is its architecture and shape, which relates to the
way branch points lead to a deviation from a simple linear chain.[28] A branched polymer molecule
is composed of a main chain with one or more substituent side chains or branches. Types of
branched polymers include star polymers, comb polymers, polymer brushes, dendronized polymers,
ladder polymers, and dendrimers.[28] There exist also two-dimensional polymers (2DP) which are
composed of topologically planar repeat units. A polymer's architecture affects many of its physical
properties including solution viscosity, melt viscosity, solubility in various solvents, glass-transition
temperature and the size of individual polymer coils in solution. A variety of techniques may be
employed for the synthesis of a polymeric material with a range of architectures, for example living
polymerization.

Chain length

A common means of expressing the length of a chain is the degree of polymerization, which
quantifies the number of monomers incorporated into the chain.[29][30] As with other molecules, a
polymer's size may also be expressed in terms of molecular weight. Since synthetic polymerization
techniques typically yield a statistical distribution of chain lengths, the molecular weight is expressed
in terms of weighted averages. The number-average molecular weight (Mn) and weight-average
molecular weight (Mw) are most commonly reported.[31][32] The ratio of these two values (Mw /
Mn) is the dispersity (Đ), which is commonly used to express the width of the molecular weight
distribution.[33]
The physical properties[34] of polymer strongly depend on the length (or equivalently, the molecular
weight) of the polymer chain.[35] One important example of the physical consequences of the
molecular weight is the scaling of the viscosity (resistance to flow) in the melt.[36] The influence of
the weight-average molecular weight ({\displaystyle M_{w}}M_{w}) on the melt viscosity
({\displaystyle \eta }\eta ) depends on whether the polymer is above or below the onset of
entanglements. Below the entanglement molecular weight[clarification needed], {\displaystyle \eta
\sim {M_{w}}^{1}}{\displaystyle \eta \sim {M_{w}}^{1}}, whereas above the entanglement molecular
weight, {\displaystyle \eta \sim {M_{w}}^{3.4}}{\displaystyle \eta \sim {M_{w}}^{3.4}}. In the latter
case, increasing the polymer chain length 10-fold would increase the viscosity over 1000
times.[37][page needed] Increasing chain length furthermore tends to decrease chain mobility,
increase strength and toughness, and increase the glass-transition temperature (Tg).[38] This is a
result of the increase in chain interactions such as van der Waals attractions and entanglements that
come with increased chain length.[39][40] These interactions tend to fix the individual chains more
strongly in position and resist deformations and matrix breakup, both at higher stresses and higher
temperatures.

Monomer arrangement in copolymers

Main article: Copolymer

Copolymers are classified either as statistical copolymers, alternating copolymers, block copolymers,
graft copolymers or gradient copolymers. In the schematic figure below, Ⓐ and Ⓑ symbolize the
two repeat units.

Statistisches Copolymer

random copolymer Gradientcopolymer

gradient copolymer Pfropfcopolymer

graft copolymer

Alternierendes Copolymer

alternating copolymer Blockcopolymer

block copolymer

Alternating copolymers possess two regularly alternating monomer residues:[41] {{not a


typo[AB]n}}. An example is the equimolar copolymer of styrene and maleic anhydride formed by
free-radical chain-growth polymerization.[42] A step-growth copolymer such as Nylon 66 can also be
considered a strictly alternating copolymer of diamine and diacid residues, but is often described as
a homopolymer with the dimeric residue of one amine and one acid as a repeat unit.[43]

Periodic copolymers have more than two species of monomer units in a regular sequence.[44]

Statistical copolymers have monomer residues arranged according to a statistical rule. A statistical
copolymer in which the probability of finding a particular type of monomer residue at a particular
point in the chain is independent of the types of surrounding monomer residue may be referred to
as a truly random copolymer.[45][46] For example, the chain-growth copolymer of vinyl chloride and
vinyl acetate is random.[42]

Block copolymers have long sequences of different monomer units.[42][43] Polymers with two or
three blocks of two distinct chemical species (e.g., A and B) are called diblock copolymers and
triblock copolymers, respectively. Polymers with three blocks, each of a different chemical species
(e.g., A, B, and C) are termed triblock terpolymers.

Graft or grafted copolymers contain side chains or branches whose repeat units have a different
composition or configuration than the main chain.[43] The branches are added on to a preformed
main chain macromolecule.[42]

Monomers within a copolymer may be organized along the backbone in a variety of ways. A
copolymer containing a controlled arrangement of monomers is called a sequence-controlled
polymer.[47] Alternating, periodic and block copolymers are simple examples of sequence-
controlled polymers.

Tacticity

Main article: Tacticity

Tacticity describes the relative stereochemistry of chiral centers in neighboring structural units
within a macromolecule. There are three types of tacticity: isotactic (all substituents on the same
side), atactic (random placement of substituents), and syndiotactic (alternating placement of
substituents).

Isotactic-A-2D-skeletal.png

isotactic Syndiotactic-2D-skeletal.png

syndiotactic Atactic-2D-skeletal.png

atactic (i. e. random)

Morphology

Polymer morphology generally describes the arrangement and microscale ordering of polymer
chains in space. The macroscopic physical properties of a polymer are related to the interactions
between the polymer chains.

Statistischer Kneul.svg

Randomly oriented polymer Verhakungen.svg

Interlocking of several polymers

Disordered polymers: In the solid state, atactic polymers, polymers with a high degree of branching
and random copolymers form amorphous (i.e. glassy structures).[48] In melt and solution, polymers
tend to form a constantly changing "statistical cluster", see freely-jointed-chain model. In the solid
state, the respective conformations of the molecules are frozen. Hooking and entanglement of chain
molecules lead to a "mechanical bond" between the chains. Intermolecular and intramolecular
attractive forces only occur at sites where molecule segments are close enough to each other. The
irregular structures of the molecules prevent a narrower arrangement.

Polyethylene-xtal-view-down-axis-3D-balls-perspective.png

polyethylene: zigzag conformation of molecules in close packed chains Lamellen.svg

lamella with tie molecules Spherulite2de.svg

spherulite

Helix-Polypropylen.svg

polypropylene helix P-Aramid H-Brücken.svg

p-Aramid, red dotted: hydrogen bonds

Linear polymers with periodic structure, low branching and stereoregularity (e. g. not atactic) have a
semi-crystalline structure in the solid state.[48] In simple polymers (such as polyethylene), the chains
are present in the crystal in zigzag conformation. Several zigzag conformations form dense chain
packs, called crystallites or lamellae. The lamellae are much thinner than the polymers are long
(often about 10 nm).[49] They are formed by more or less regular folding of one or more molecular
chains. Amorphous structures exist between the lamellae. Individual molecules can lead to
entanglements between the lamellae and can also be involved in the formation of two (or more)
lamellae (chains than called tie molecules). Several lamellae form a superstructure, a spherulite,
often with a diameter in the range of 0.05 to 1 mm.[49]

The type and arrangement of (functional) residues of the repeat units effects or determines the
crystallinity and strength of the secondary valence bonds. In isotactic polypropylene, the molecules
form a helix. Like the zigzag conformation, such helices allow a dense chain packing. Particularly
strong intermolecular interactions occur when the residues of the repeating units allow the
formation of hydrogen bonds, as in the case of p-aramid. Crystallinity and superstructure are always
dependent on the conditions of their formation, see also: crystallization of polymers. Compared to
amorphous structures, semi-crystalline structures lead to a higher stiffness, density, melting
temperature and higher resistance of a polymer.

Cross-linked polymers: Wide-meshed cross-linked polymers are elastomers and cannot be molten
(unlike thermoplastics); heating cross-linked polymers only leads to decomposition. Thermoplastic
elastomers, on the other hand, are reversibly "physically crosslinked" and can be molten. Block
copolymers in which a hard segment of the polymer has a tendency to crystallize and a soft segment
has an amorphous structure are one type of thermoplastic elastomers: the hard segments ensure
wide-meshed, physical crosslinking.

Polymerstruktur-weitmaschig vernetzt.svg

wide-meshed cross-linked polymer (elastomer) Polymerstruktur-weitmaschig vernetzt-gestreckt.svg

wide-meshed cross-linked polymer (elastomer) under tensile stress Polymerstruktur-TPE-


teilkristallin.svg
crystallites as "crosslinking sites": one type of thermoplastic elastomer Polymerstruktur-TPE-
teilkristallin gestreckt.svg

semi-crystalline thermoplastic elastomer under tensile stress

Crystallinity

When applied to polymers, the term crystalline has a somewhat ambiguous usage. In some cases,
the term crystalline finds identical usage to that used in conventional crystallography. For example,
the structure of a crystalline protein or polynucleotide, such as a sample prepared for x-ray
crystallography, may be defined in terms of a conventional unit cell composed of one or more
polymer molecules with cell dimensions of hundreds of angstroms or more. A synthetic polymer may
be loosely described as crystalline if it contains regions of three-dimensional ordering on atomic
(rather than macromolecular) length scales, usually arising from intramolecular folding or stacking of
adjacent chains. Synthetic polymers may consist of both crystalline and amorphous regions; the
degree of crystallinity may be expressed in terms of a weight fraction or volume fraction of
crystalline material. Few synthetic polymers are entirely crystalline.[50] The crystallinity of polymers
is characterized by their degree of crystallinity, ranging from zero for a completely non-crystalline
polymer to one for a theoretical completely crystalline polymer. Polymers with microcrystalline
regions are generally tougher (can be bent more without breaking) and more impact-resistant than
totally amorphous polymers.[51] Polymers with a degree of crystallinity approaching zero or one will
tend to be transparent, while polymers with intermediate degrees of crystallinity will tend to be
opaque due to light scattering by crystalline or glassy regions. For many polymers, reduced
crystallinity may also be associated with increased transparency.

Chain conformation

The space occupied by a polymer molecule is generally expressed in terms of radius of gyration,
which is an average distance from the center of mass of the chain to the chain itself. Alternatively, it
may be expressed in terms of pervaded volume, which is the volume spanned by the polymer chain
and scales with the cube of the radius of gyration.[52] The simplest theoretical models for polymers
in the molten, amorphous state are ideal chains.

Properties

Polymer properties depend of their structure and they are divided into classes according to their
physical basis. Many physical and chemical properties describe how a polymer behaves as a
continuous macroscopic material. They are classified as bulk properties, or intensive properties
according to thermodynamics.

Mechanical properties

A polyethylene sample that has necked under tension.


The bulk properties of a polymer are those most often of end-use interest. These are the properties
that dictate how the polymer actually behaves on a macroscopic scale.

Tensile strength

The tensile strength of a material quantifies how much elongating stress the material will endure
before failure.[53][54] This is very important in applications that rely upon a polymer's physical
strength or durability. For example, a rubber band with a higher tensile strength will hold a greater
weight before snapping. In general, tensile strength increases with polymer chain length and
crosslinking of polymer chains.

Young's modulus of elasticity

Young's modulus quantifies the elasticity of the polymer. It is defined, for small strains, as the ratio
of rate of change of stress to strain. Like tensile strength, this is highly relevant in polymer
applications involving the physical properties of polymers, such as rubber bands. The modulus is
strongly dependent on temperature. Viscoelasticity describes a complex time-dependent elastic
response, which will exhibit hysteresis in the stress-strain curve when the load is removed. Dynamic
mechanical analysis or DMA measures this complex modulus by oscillating the load and measuring
the resulting strain as a function of time.

Transport properties

Transport properties such as diffusivity describe how rapidly molecules move through the polymer
matrix. These are very important in many applications of polymers for films and membranes.

The movement of individual macromolecules occurs by a process called reptation in which each
chain molecule is constrained by entanglements with neighboring chains to move within a virtual
tube. The theory of reptation can explain polymer molecule dynamics and viscoelasticity.[55]

Phase behavior

Crystallization and melting

Thermal transitions in (A) amorphous and (B) semicrystalline polymers, represented as traces from
differential scanning calorimetry. As the temperature increases, both amorphous and semicrystalline
polymers go through the glass transition (Tg). Amorphous polymers (A) do not exhibit other phase
transitions, though semicrystalline polymers (B) undergo crystallization and melting (at
temperatures Tc and Tm, respectively).

Depending on their chemical structures, polymers may be either semi-crystalline or amorphous.


Semi-crystalline polymers can undergo crystallization and melting transitions, whereas amorphous
polymers do not. In polymers, crystallization and melting do not suggest solid-liquid phase
transitions, as in the case of water or other molecular fluids. Instead, crystallization and melting
refer to the phase transitions between two solid states (i.e., semi-crystalline and amorphous).
Crystallization occurs above the glass-transition temperature (Tg) and below the melting
temperature (Tm).

Glass transition

All polymers (amorphous or semi-crystalline) go through glass transitions. The glass-transition


temperature (Tg) is a crucial physical parameter for polymer manufacturing, processing, and use.
Below Tg, molecular motions are frozen and polymers are brittle and glassy. Above Tg, molecular
motions are activated and polymers are rubbery and viscous. The glass-transition temperature may
be engineered by altering the degree of branching or crosslinking in the polymer or by the addition
of plasticizers.[56]

Whereas crystallization and melting are first-order phase transitions, the glass transition is not.[57]
The glass transition shares features of second-order phase transitions (such as discontinuity in the
heat capacity, as shown in the figure), but it is generally not considered a thermodynamic transition
between equilibrium states.

Mixing behavior

Phase diagram of the typical mixing behavior of weakly interacting polymer solutions, showing
spinodal curves and binodal coexistence curves.

In general, polymeric mixtures are far less miscible than mixtures of small molecule materials. This
effect results from the fact that the driving force for mixing is usually entropy, not interaction
energy. In other words, miscible materials usually form a solution not because their interaction with
each other is more favorable than their self-interaction, but because of an increase in entropy and
hence free energy associated with increasing the amount of volume available to each component.
This increase in entropy scales with the number of particles (or moles) being mixed. Since polymeric
molecules are much larger and hence generally have much higher specific volumes than small
molecules, the number of molecules involved in a polymeric mixture is far smaller than the number
in a small molecule mixture of equal volume. The energetics of mixing, on the other hand, is
comparable on a per volume basis for polymeric and small molecule mixtures. This tends to increase
the free energy of mixing for polymer solutions and thereby making solvation less favorable, and
thereby making the availability of concentrated solutions of polymers far rarer than those of small
molecules.

Furthermore, the phase behavior of polymer solutions and mixtures is more complex than that of
small molecule mixtures. Whereas most small molecule solutions exhibit only an upper critical
solution temperature phase transition (UCST), at which phase separation occurs with cooling,
polymer mixtures commonly exhibit a lower critical solution temperature phase transition (LCST), at
which phase separation occurs with heating.
In dilute solutions, the properties of the polymer are characterized by the interaction between the
solvent and the polymer. In a good solvent, the polymer appears swollen and occupies a large
volume. In this scenario, intermolecular forces between the solvent and monomer subunits
dominate over intramolecular interactions. In a bad solvent or poor solvent, intramolecular forces
dominate and the chain contracts. In the theta solvent, or the state of the polymer solution where
the value of the second virial coefficient becomes 0, the intermolecular polymer-solvent repulsion
balances exactly the intramolecular monomer-monomer attraction. Under the theta condition (also
called the Flory condition), the polymer behaves like an ideal random coil. The transition between
the states is known as a coil–globule transition.

Inclusion of plasticizers

Inclusion of plasticizers tends to lower Tg and increase polymer flexibility. Addition of the plasticizer
will also modify dependence of the glass-transition temperature Tg on the cooling rate.[58] The
mobility of the chain can further change if the molecules of plasticizer give rise to hydrogen bonding
formation. Plasticizers are generally small molecules that are chemically similar to the polymer and
create gaps between polymer chains for greater mobility and reduced interchain interactions. A
good example of the action of plasticizers is related to polyvinylchlorides or PVCs. A uPVC, or
unplasticized polyvinylchloride, is used for things such as pipes. A pipe has no plasticizers in it,
because it needs to remain strong and heat-resistant. Plasticized PVC is used in clothing for a flexible
quality. Plasticizers are also put in some types of cling film to make the polymer more flexible.

Chemical properties

The attractive forces between polymer chains play a large part in determining the polymer’s
properties. Because polymer chains are so long, they have many such interchain interactions per
molecule, amplifying the effect of these interactions on the polymer properties in comparison to
attractions between conventional molecules. Different side groups on the polymer can lend the
polymer to ionic bonding or hydrogen bonding between its own chains. These stronger forces
typically result in higher tensile strength and higher crystalline melting points.

The intermolecular forces in polymers can be affected by dipoles in the monomer units. Polymers
containing amide or carbonyl groups can form hydrogen bonds between adjacent chains; the
partially positively charged hydrogen atoms in N-H groups of one chain are strongly attracted to the
partially negatively charged oxygen atoms in C=O groups on another. These strong hydrogen bonds,
for example, result in the high tensile strength and melting point of polymers containing urethane or
urea linkages. Polyesters have dipole-dipole bonding between the oxygen atoms in C=O groups and
the hydrogen atoms in H-C groups. Dipole bonding is not as strong as hydrogen bonding, so a
polyester's melting point and strength are lower than Kevlar's (Twaron), but polyesters have greater
flexibility. Polymers with non-polar units such as polyethylene interact only through weak Van der
Waals forces. As a result, they typically have lower melting temperatures than other polymers.
When a polymer is dispersed or dissolved in a liquid, such as in commercial products like paints and
glues, the chemical properties and molecular interactions influence how the solution flows and can
even lead to self-assembly of the polymer into complex structures. When a polymer is applied as a
coating, the chemical properties will influence the adhesion of the coating and how it interacts with
external materials, such as superhydrophobic polymer coatings leading to water resistance. Overall
the chemical properties of a polymer are important elements for designing new polymeric material
products.

Optical properties

Polymers such as PMMA and HEMA:MMA are used as matrices in the gain medium of solid-state dye
lasers, also known as solid-state dye-doped polymer lasers. These polymers have a high surface
quality and are also highly transparent so that the laser properties are dominated by the laser dye
used to dope the polymer matrix. These type of lasers, that also belong to the class of organic lasers,
are known to yield very narrow linewidths which is useful for spectroscopy and analytical
applications.[59] An important optical parameter in the polymer used in laser applications is the
change in refractive index with temperature also known as dn/dT. For the polymers mentioned here
the (dn/dT) ~ −1.4 × 10−4 in units of K−1 in the 297 ≤ T ≤ 337 K range.[60]

Electrical properties

Most conventional polymers such as polyethylene are electrical insulators, but the development of
polymers containing π-conjugated bonds has led to a wealth of polymer-based semiconductors, such
as polythiophenes. This has led to many applications in the field of organic electronics.

Applications

Nowadays, synthetic polymers are used in almost all walks of life. Modern society would look very
different without them. The spreading of polymer use is connected to their unique properties: low
density, low cost, good thermal/electrical insulation properties, high resistance to corrosion, low-
energy demanding polymer manufacture and facile processing into final products. For a given
application, the properties of a polymer can be tuned or enhanced by combination with other
materials, as in composites. Their application allows to save energy (lighter cars and planes,
thermally insulated buildings), protect food and drinking water (packaging), save land and reduce
use of fertilizers (synthetic fibres), preserve other materials (coatings), protect and save lifes
(hygiene, medical applications). A representative, non-exhaustive list of applications is given below.

Clothing, sportswear and accessories: polyester and PVC clothing, spandex, sport shoes, wetsuits,
footballs and billiard balls, skis and snowboards, rackets, parachutes, sails, tents and shelters.

Electronic and photonic technologies: organic field effect transistors (OFET), light emitting diodes
(OLED) and solar cells, television components, compact discs (CD), photoresists, holography.

Packaging and containers: films, bottles, food packaging, barrels.

Insulation: electrical and thermal insulation, spray foams.


Construction and structural applications: garden furniture, PVC windows, flooring, sealing, pipes.

Paints, glues and lubricants: varnish, adhesives, dispersants, anti-graffiti coatings, antifouling
coatings, non-stick surfaces, lubricants.

Car parts: tires, bumpers, windshields, windscreen wipers, fuel tanks, car seats.

Household items: buckets, kitchenware, toys (e.g., construction sets and Rubik's cube).

Medical applications: blood bag, syringes, rubber gloves, surgical suture, contact lenses, prosthesis,
controlled drug delivery and release, matrices for cell growth.

Personal hygiene and healthcare: diapers using superabsorbent polymers, toothbrushes, cosmetics,
shampoo, condoms.

Security: personal protective equipment, bulletproof vests, space suits, ropes.

Separation technologies: synthetic membranes, fuel cell membranes, filtration, ion-exchange resins.

Money: polymer banknotes and payment cards.

3D printing.

Standardized nomencla

You might also like