Fact-checked by Grok 2 weeks ago

Polymer

A polymer is a substance composed of macromolecules, which are very large molecules with molecular weights ranging from a few thousand to millions of grams per mole, formed by the multiple repetition of smaller units derived from molecules of low relative molecular mass. These macromolecules consist of repeating structural units known as monomers, which link together through chemical bonds to create long chains, often with branching or cross-linking that influences the material's properties. The term "polymer" was coined by German chemist in the 1920s, who established the macromolecular hypothesis and received the in 1953 for his foundational work. Polymers can be classified as natural or synthetic, with natural polymers occurring in living organisms and synthetic ones produced through industrial processes. polymers include proteins, , , and DNA, which have been utilized by humans since prehistoric times for applications such as textiles and . Synthetic polymers, first developed in the early 20th century, encompass materials like phenol-formaldehyde (, invented around 1905–1909), poly(vinyl chloride) (PVC, commercialized in 1930), (PET, 1943), and . They are synthesized via two primary mechanisms: addition , where monomers link without loss of atoms (e.g., from ), and condensation , which involves the elimination of small molecules like (e.g., from diamines and diacids). Homopolymers consist of identical repeating units, while copolymers incorporate two or more different monomers for tailored properties. The properties of polymers vary widely depending on their structure, molecular weight, and processing, enabling diverse applications across industries. Global production of polymers exceeded 450 million metric tons in 2024. For instance, (HDPE) is rigid with a of 130°C, while (LDPE) is softer at 110°C; elastomers like rubber exhibit high elasticity with melting points around 30°C. Polymers are integral to daily life, forming plastics, fibers, elastomers, adhesives, and composites used in , , medical devices (e.g., heart valves), transportation, , and . They drive innovations in , , and sustainable materials.

Etymology and History

Etymology

The term "polymer" originates from the Greek words poly (πολύς), meaning "many," and meros (μέρος), meaning "parts," and was coined in by the to describe compounds sharing the same empirical composition but differing in molecular weight by integral multiples, such as and butylene. introduced this terminology in the context of to denote a specific type of isomerism, without implying the long-chain structures understood today. In 1861, British chemist Thomas Graham extended the term "polymer" to describe colloidal substances, proposing that materials like and were aggregates or "polymeric" associations of smaller molecules held together by weak forces, as part of his association theory contrasting colloids with crystalloids. This usage marked an early application to high-molecular-weight substances exhibiting low , laying groundwork for later interpretations in macromolecular chemistry. The concept evolved significantly in the early 20th century through the work of Hermann Staudinger, who in the 1920s advocated for the macromolecular hypothesis, redefining polymers as long-chain molecules formed by covalent linkages of monomeric units rather than mere aggregates. Staudinger coined the term "macromolecule" (Makromolekül) in 1922 to emphasize the enormous size of these structures, distinguishing them from Berzelius's original compositional sense and Graham's colloidal view. These developments solidified "polymer" in its modern sense, focusing on chain-like macromolecules central to both natural and synthetic materials.

Historical Development

The utilization of natural polymers dates back to ancient civilizations. In , around 2700 BCE, the production of —a protein-based polymer derived from silkworm cocoons—emerged as a key technological achievement, enabling the weaving of fine fabrics that became central to and . Similarly, in , by 1600 BCE, indigenous peoples such as the Olmec and processed latex from the tree, mixing it with morning glory vine juice to create solid rubber for balls, seals, and other tools, demonstrating early mastery of natural polymer manipulation. The 19th century marked the transition toward synthetic polymers through industrial innovations. In 1839, American inventor discovered , a process that heated with to enhance its elasticity and durability, revolutionizing its commercial viability for tires and . Two decades later, in 1862, British chemist patented Parkesine, the first man-made plastic derived from cellulose nitrate, which could be molded into durable items like combs and buttons, laying the groundwork for the . In 1907, Belgian-American chemist invented , the first fully synthetic plastic, through the condensation of phenol and , initiating the era of commercial thermosetting plastics. The 20th century saw foundational scientific breakthroughs that established as a distinct field. In 1920, German chemist proposed the macromolecular hypothesis, arguing that polymers consist of long chains of covalently bonded monomers rather than aggregates of small molecules, a concept validated over decades and earning him the in 1953. Building on this, in 1935, American chemist at synthesized , the first fully synthetic fiber, by polycondensing and , enabling mass production of strong, versatile materials for textiles and more. In the 1950s, and developed Ziegler-Natta catalysts, enabling stereospecific polymerization of olefins like and into and isotactic , innovations recognized with the 1963 . Entering the 21st century, polymer research shifted toward functional and sustainable materials. In 2000, the was awarded to , Alan G. MacDiarmid, and Hideki Shirakawa for discovering conductive polymers, such as doped , which conduct electricity like metals while retaining polymer flexibility, opening applications in electronics and sensors. In the 2020s, attention has intensified on biodegradable polymers like (PHA), microbial polyesters that fully degrade in natural environments, addressing plastic waste through sustainable alternatives in packaging and agriculture. By 2025, advancements in bio-based polymers from renewable feedstocks, such as sugarcane-derived and CO2-captured materials, have accelerated, with companies like introducing innovations for reusable packaging and construction, driven by demands.

Classification and Examples

Natural Polymers

Natural polymers, also known as biopolymers, are large molecules synthesized by living organisms through enzymatic processes, consisting of repeating monomeric units covalently linked to form chains or networks. These include three primary classes: proteins, formed from amino acid monomers; nucleic acids, composed of nucleotide units such as in DNA and RNA; and polysaccharides, built from sugar monomers. Unlike synthetic polymers, biopolymers are produced in biological systems and play essential roles in cellular structure, function, and metabolism. Prominent examples derive from diverse biological sources, with plants serving as the primary origin for many abundant polysaccharides. Cellulose, a linear polysaccharide of glucose units linked by β-1,4-glycosidic bonds, is the most prevalent organic polymer on Earth, accounting for approximately 33% of all plant biomass and forming the structural framework of plant cell walls. Starch, another plant-derived polysaccharide composed of α-glucose units, functions mainly in energy storage in seeds, roots, and tubers. In animals, proteins such as collagen predominate; collagen is a fibrous protein assembled from glycine, proline, and hydroxyproline-rich sequences, forming a triple helix that provides tensile strength in connective tissues like skin, tendons, and bones. Nucleic acids originate from all living cells, with DNA serving as the genetic blueprint in nuclei and RNA facilitating protein synthesis. Chitin, a polysaccharide of N-acetylglucosamine, forms the exoskeletons of arthropods and fungal cell walls. Natural rubber, a polyisoprene elastomer, is extracted from the latex of the Hevea brasiliensis tree, where it exists as colloidal particles in the sap. Lignin, a complex aromatic heteropolymer derived from phenylpropanoid units, impregnates plant cell walls, particularly in wood, contributing up to one-third of its dry weight. In nature, these polymers fulfill critical structural, storage, and informational roles. Cellulose and lignin provide mechanical support and rigidity to plant tissues, enabling upright growth and resistance to environmental stresses. and (an analog) act as energy reserves, broken down into glucose during metabolic needs. Proteins like maintain tissue integrity and elasticity in animal extracellular matrices, while enzymes (also proteins) catalyze biochemical reactions. Nucleic acids store and transmit hereditary information, with DNA's double-helix ensuring stable replication and RNA enabling . offers protective barriers in invertebrates and fungi, and in may deter herbivores or seal wounds. These functions underscore the evolutionary adaptation of to sustain life processes across kingdoms.

Synthetic Polymers

Synthetic polymers are human-made materials produced through in laboratories, typically derived from petroleum-based or bio-based monomers such as or . Unlike natural polymers like or proteins, which occur in biological systems, synthetic polymers offer greater versatility in structure and properties due to controlled processes. These polymers are broadly classified into major categories based on their thermal and mechanical behaviors: thermoplastics, thermosets, and elastomers. Thermoplastics, which soften upon heating and can be reshaped repeatedly, include and (PVC); for instance, (HDPE) and (LDPE) are widely used in packaging due to their durability and flexibility. Thermosets, such as epoxy resins, form irreversible cross-linked networks during curing, resulting in rigid structures with high thermal stability suitable for adhesives and composites. Elastomers, characterized by high elasticity and resilience, encompass synthetic rubbers like styrene-butadiene rubber (SBR), which mimics the properties of but offers improved resistance to abrasion and aging. Prominent examples highlight the diversity of synthetic polymers. serves as a lightweight foam material for insulation and packaging, valued for its low cost and ease of molding. Polyurethanes, formed from diisocyanates and polyols, are employed in flexible foams for cushions and durable coatings for surfaces, providing a balance of toughness and elasticity. A notable bio-derived addition is (PLA), an aliphatic polyester produced from renewable sources like , which was commercialized in the early 1990s and is prized for its biodegradability and use in packaging and medical applications. The design of synthetic polymers emphasizes tailoring molecular to achieve targeted , such as enhanced through cross-linking or flexibility via linear chain structures. By selecting specific monomers, adjusting molecular weight, and controlling conditions, engineers customize these materials for applications requiring precise mechanical, thermal, or chemical performance. This intentional engineering distinguishes synthetic polymers, enabling innovations beyond the limitations of natural counterparts.

Molecular Structure

Monomers and Repeat Units

Polymers are formed from small molecules known as , which are capable of linking together through chemical to create long chains or networks. A typically contains functional groups that enable , such as double bonds in or reactive end groups in bifunctional molecules. For instance, (C₂H₄), a simple , serves as the for , one of the most common synthetic polymers. In biological systems, act as , linking to form proteins; each has an amino group and a carboxyl group that participate in bond formation. Upon , are incorporated into the polymer chain, resulting in a repeating structural segment called the constitutional repeating unit (CRU), which is the smallest identifiable repeating portion of the polymer backbone. The CRU is determined by examining the polymer's connectivity and selecting the subunit that, when repeated, reconstructs the chain with the lowest possible locants for substituents. For , the CRU is -[\ce{CH2-CH2}]-, derived directly from the monomer after opening its . In proteins, the CRU consists of the amide-linked backbone from , excluding the variable side chains. The linkages between monomers occur via covalent bonds formed in two primary mechanisms: polyaddition and polycondensation. In polyaddition, react without eliminating small molecules, directly incorporating the entire monomer structure into the ; this is common for monomers with carbon-carbon double bonds, as in the formation of from . In polycondensation, link with the release of a small , such as , resulting in a that may differ slightly from the original ; for example, form bonds in proteins by eliminating H₂O from the carboxyl and amino groups. The general representation of is n \, M \rightarrow [M]_n, where M denotes the and n is the , indicating the number of repeat units in the chain. Polymers are classified as homopolymers or copolymers based on the number of distinct types. Homopolymers consist of a single repeating type, such as derived solely from , leading to a uniform CRU throughout the chain. Copolymers, in contrast, incorporate two or more different s, resulting in sequences of varied repeat units; often uses connectives like "co-" to denote this, as in poly(styrene-co-butadiene). This distinction allows for tailored properties in materials design.

Microstructure

The microstructure of polymers refers to the arrangement and configuration of monomer units within the polymer chains, which significantly influences their physical and chemical behavior. Polymer architecture encompasses various structural motifs, including linear, branched, cross-linked, star, and dendrimer forms. In linear polymers, monomer units connect in a straight chain without side branches, as seen in . Branched architectures feature side chains attached to the main backbone, such as in produced via free-radical , where short-chain branches arise from intramolecular hydrogen abstraction during . Cross-linked polymers involve covalent bonds between different chains, forming networks that enhance rigidity, while star polymers consist of multiple linear arms radiating from a central , and dendrimers exhibit highly ordered, tree-like branching with precise generational layers. These architectures are tailored through synthesis methods to achieve desired properties, with branching generally increasing chain entanglement and altering flow characteristics. Chain length in polymers is quantified by molecular weight metrics, reflecting the . The number-average molecular weight (M_n) is the of the molecular weights of all chains, calculated as the total divided by the number of molecules, while the weight-average molecular weight (M_w) weights each chain by its , emphasizing longer chains and typically yielding higher values than M_n. The polydispersity (PDI), defined as: \text{PDI} = \frac{M_w}{M_n} measures the breadth of the molecular ; a PDI of 1 indicates monodispersity (uniform chain lengths), but most synthetic polymers have PDI > 1, signifying a of lengths that broadens with less controlled , thereby increasing melt and processing challenges. These parameters are determined experimentally via techniques like . Copolymers, formed from two or more distinct , exhibit varied microstructures based on monomer sequencing. Random copolymers have monomers distributed irregularly along the chain, leading to averaged properties; alternating copolymers feature strict ABAB patterns, often due to charge-transfer interactions in copolymerization; block copolymers consist of long sequences of one monomer type followed by another (e.g., AAAAABBBB), enabling into domains; and graft copolymers attach branches of one monomer type onto a backbone of another. For instance, block copolymers can self-assemble into ordered structures like micelles in selective solvents due to incompatible blocks. These configurations are controlled by polymerization techniques such as living anionic polymerization for blocks. Tacticity describes the stereochemical arrangement of substituents along the polymer backbone in polymers, arising from the at each . Isotactic polymers have all substituents on the same side of the chain (regular configuration), syndiotactic polymers alternate sides, and atactic polymers show random placement, resulting in amorphous structures. Stereoregular isotactic and syndiotactic polymers, which enable higher order, are synthesized using Ziegler-Natta catalysts—heterogeneous systems of compounds (e.g., TiCl₄) and organoaluminum cocatalysts—that coordinate monomers in a specific during , as pioneered in the for polypropene production. This stereocontrol revolutionized synthesis, allowing crystalline materials with enhanced strength.

Morphology

Polymer morphology refers to the physical arrangement and organization of polymer chains in bulk materials, which determines many macroscopic properties such as mechanical strength and optical clarity. In amorphous regions, polymer chains typically adopt conformations, characterized by disordered, entangled structures that maximize , as described in Flory's of real polymer chains. In contrast, within crystalline domains, chains assume more ordered conformations, such as extended planar zigzags in or helical arrangements in isotactic polymers like , enabling close packing and higher density. Crystallinity represents the degree of structural order in these crystalline domains, often quantified as the percentage of crystalline material relative to the total mass, with typical values ranging from 50% to 90% in (HDPE). This degree is commonly measured using (DSC), where the heat of fusion is compared to that of a fully crystalline . Spherulites serve as the primary growth units in semicrystalline polymers, forming radially branching aggregates of lamellar crystals from a central , as explained by the phenomenological theory of and Padden, which attributes their development to the of noncrystallizing material ahead of the front. Semicrystalline polymers consist of alternating crystalline and amorphous regions, while fully amorphous polymers lack long-range order. In the amorphous components of both types, the material exists in a glassy below the temperature (), where chains are rigid and immobile due to restricted segmental motion, transitioning to a rubbery above with increased chain flexibility and elasticity. The degree of crystallinity is also influenced by the of the polymer chains, as detailed in the microstructure section. Morphology is significantly affected by processing conditions, such as cooling rate during solidification. For instance, rapid quenching of poly(ethylene terephthalate) () at rates of 1 or higher yields a fully amorphous structure by preventing chain reorganization into , whereas slower cooling promotes partial or full .

Synthesis

Polymerization Mechanisms

Polymerization mechanisms encompass the chemical pathways through which monomers link to form long-chain synthetic polymers, primarily classified into chain-growth () and step-growth types. These mechanisms differ fundamentally in how molecular weight develops and the nature of reactive intermediates involved. Chain-growth polymerization proceeds via sequential to active chain ends, enabling rapid molecular weight buildup even at low monomer conversion, while step-growth relies on intermolecular reactions between functional groups, requiring high conversion for substantial chain lengths. Coordination mechanisms, a subset of chain-growth, utilize metal catalysts for precise control over polymer . Addition polymerization, or , involves the opening of double bonds in vinyl or similar monomers through reactive species like free radicals, carbocations, carbanions, or metal complexes. In free radical addition polymerization, begins with the or photochemical decomposition of an initiator, such as a , generating radicals that add to the monomer's , forming a chain-carrying . Propagation continues as this radical adds successive monomers, exemplified by the of to (PVC), a widely used . Termination occurs via radical combination or , limiting chain length and broadening the molecular weight distribution. This mechanism, kinetically described by Flory in , dominates industrial production of polymers like and due to its simplicity and tolerance for impurities. Ionic variants of addition polymerization include cationic and anionic mechanisms, which offer greater control over chain architecture. Cationic polymerization employs electrophilic initiators like Lewis acids to generate carbocations, suitable for monomers such as , but often limited by . Anionic polymerization, conversely, uses nucleophilic initiators like alkyllithium compounds, propagating via carbanions. The seminal discovery of living anionic polymerization by Szwarc in 1956 demonstrated that, in the absence of terminating impurities, chains remain active, allowing precise molecular weight control and narrow polydispersity index (PDI, typically <1.1). This enables synthesis of block copolymers, as seen in styrene-butadiene-styrene triblock polymers for thermoplastic elastomers. Living techniques have since extended to cationic systems using weakly coordinating counterions, further expanding access to tailored architectures. A major advancement in addition polymerization is controlled radical polymerization (CRP), which achieves living-like characteristics in free radical systems through reversible deactivation of chain-end radicals. Key methods include (ATRP), developed by Matyjaszewski and Sawamoto in 1995, which uses catalysts (e.g., complexes) to reversibly oxidize radicals to dormant alkyl halides; reversible addition-fragmentation (RAFT), introduced by CSIRO researchers in 1998, employing thiocarbonylthio compounds as agents; and nitroxide-mediated polymerization (NMP), pioneered by Georges et al. in 1993, utilizing stable nitroxide radicals for reversible trapping. These techniques yield polymers with predetermined molecular weights, low PDI (typically <1.5), and high chain-end fidelity, facilitating the synthesis of like stimuli-responsive hydrogels and nanostructured films. CRP's versatility with a wide range of monomers and tolerance to functional groups has made it indispensable in academic and industrial settings since the late . Step-growth polymerization forms polymers through repeated reactions between bifunctional monomers, often via condensation with elimination of small byproducts like water. Unlike chain-growth, active species are the functional groups themselves, leading to formation first, followed by gradual chain extension. A classic example is the synthesis of nylon 6,6, a produced by the condensation of and , where bonds form and water is released. This process, pioneered by Carothers at in the 1930s, requires stoichiometric balance and high purity to achieve high molecular weights. The relationship between chain length and reaction progress is quantified by the : X_n = \frac{1}{1 - p} where X_n is the number-average degree of polymerization and p is the extent of reaction (fraction of functional groups consumed). For instance, at p = 0.99, X_n \approx 100, illustrating the need for near-complete conversion. Coordination polymerization, another chain-growth variant, employs transition metal catalysts to coordinate and insert monomers into a growing chain, enabling stereoregular polymers. The , developed independently by and Natta in the early , uses compounds (e.g., TiCl₄) with aluminum alkyls to polymerize α-olefins like into isotactic , a crystalline with superior mechanical properties. The mechanism involves migratory insertion at the metal center, with the catalyst's active sites dictating tacticity via monomer approach geometry. polymerization, utilizing or catalysts, extends this control to cyclic monomers, forming polymers with defined microstructures, though Ziegler-Natta remains dominant for polyolefins due to its scalability. Kinetically, chain-growth and step-growth mechanisms contrast sharply in molecular weight evolution. In chain-growth, such as free radical or coordination processes, high molecular weights emerge rapidly after , with proportional to the ratio of to termination rates, often yielding PDI around 1.5–2. In step-growth, molecular weight increases gradually, following the , where low p (e.g., 0.95) limits X_n to about 20, necessitating advanced techniques like for equilibrium shifts. These differences influence reactor design and product uniformity, with chain-growth favoring continuous processes and step-growth batch reactions. Biological polymerization variants, such as enzymatic chain-growth, mirror these principles but occur .

Biological Synthesis

Biological synthesis of polymers occurs in living organisms through highly regulated enzymatic and metabolic processes that ensure precise control over chain length, structure, and functionality. These pathways leverage cellular machinery to assemble macromolecules from simple monomers, often integrating energy from metabolic intermediates and responding to environmental cues. Unlike synthetic methods, biological polymerization emphasizes , folding, and integration into cellular functions, such as , , and . Proteins, linear polymers of , are synthesized via on ribosomes, where (mRNA) templates direct the assembly. During , transfer RNAs (tRNAs) deliver activated to the ribosome's center, catalyzing the formation of bonds between the carboxyl group of the growing chain and the amino group of the incoming . This process adds sequentially, yielding polypeptides that fold into functional proteins, with ribosomes ensuring fidelity through codon-anticodon matching and mechanisms. Polysaccharides, such as and , are built through enzymatic pathways that extend glycosidic bonds from activated sugar nucleotides like UDP-glucose. , a key , polymerizes glucose units via α-1,4 linkages to form linear chains, while branching enzymes introduce α-1,6 branches, enhancing and accessibility for rapid mobilization as reserves. These reactions occur in the of eukaryotic cells or bacterial , with regulatory modulating activity to balance synthesis and degradation. Nucleic acids, including DNA and RNA, are polymerized by nucleotidyl transferases that add nucleotides to a growing strand. DNA polymerase catalyzes the formation of phosphodiester bonds by incorporating deoxynucleoside triphosphates (dNTPs) complementary to the template strand, exclusively in the 5' to 3' direction, using the 3'-hydroxyl group of the last nucleotide as the nucleophile. This directionality ensures efficient replication with high fidelity, aided by proofreading exonuclease activity that removes mismatches, achieving error rates as low as 10^{-9} per base pair. RNA polymerase follows a similar mechanism for transcription, producing RNA strands that serve as templates or functional molecules. A notable example of bacterial is the production of (PHAs), biodegradable polyesters accumulated as carbon storage granules. In organisms like , the pathway begins with derived from β-oxidation or sugar , which is converted to 3-hydroxyacyl-CoA intermediates by β-ketothiolase and acetoacetyl-CoA reductase. PHA then polymerizes these monomers into granules, with chain lengths varying from short (PHB) to medium (), enabling applications as eco-friendly plastics that degrade in soil within months. This process is upregulated under nutrient limitation, highlighting metabolic flexibility in prokaryotes.

Modification of Natural Polymers

Modification of natural polymers involves chemical alterations to their structures, such as derivatization and cross-linking, to improve properties like , strength, and for applications. These techniques transform inherently limited natural materials into versatile derivatives while retaining core and renewability. Derivatization replaces or adds functional groups to the polymer backbone, enhancing reactivity or processability, whereas cross-linking forms covalent bonds between chains to increase rigidity and . Such modifications have enabled widespread use in textiles, adhesives, and biomedical materials, bridging natural and synthetic polymer domains. A prominent derivatization example is the conversion of cellulose to rayon via the viscose process, which begins with xanthation. In this method, alkali cellulose reacts with carbon disulfide to form cellulose xanthate, a soluble intermediate that is extruded into an acid bath to regenerate cellulose filaments with improved flexibility and dyeability compared to native cellulose. Discovered by Charles Frederick Cross and Edward John Bevan in 1891, this process marked a pivotal advancement in textile production, allowing natural cellulose from wood pulp to yield synthetic-like fibers. Cross-linking exemplifies another key technique, notably in the vulcanization of () with . Heating rubber with 1-3% creates bridges between polymer chains, transforming the sticky, temperature-sensitive material into a resilient resistant to abrasion and environmental degradation. This process, invented by in 1839, revolutionized and seal manufacturing by enhancing elasticity and longevity. Grafting copolymerization further expands modification options by attaching synthetic polymer chains to natural backbones, yielding hybrid materials with tailored properties. For instance, free-radical initiation attaches acrylic monomers to like or , combining the biodegradability of the natural component with the hydrophilicity or strength of synthetics. This approach minimizes and enables applications in and composites. Specific examples illustrate practical outcomes. with via ceric initiation produces superabsorbent polymers that swell up to 500 times their weight in water, used in products for superior absorbency over pure synthetics due to enhanced . Similarly, deacetylation of —extracted from crustacean shells—yields , a cationic polymer with 70-95% deacetylation degree, imparting antimicrobial activity against bacteria like by disrupting cell membranes. This modification improves solubility in acidic media and enables uses in dressings and . Nitrocellulose, derived from nitration of cellulose with nitric and sulfuric acids, exemplifies early derivatization for high-impact applications. Discovered by Christian Friedrich Schönbein in 1846, this material features nitrate ester groups that confer flammability and in organic solvents, leading to its use in propellants and lacquers with velocities exceeding 6000 m/s. Overall, these modifications enhance natural polymers' and , facilitating their integration into modern materials while preserving sustainability.

Properties

Mechanical Properties

Mechanical properties of polymers describe their response to applied forces, encompassing behaviors such as deformation, strength, and recovery under . These properties are crucial for determining suitability in applications ranging from structural components to flexible materials. Polymers exhibit a wide range of mechanical behaviors depending on their and processing, often falling into categories like brittle, ductile, or . Tensile strength represents the maximum a polymer can withstand while being stretched before it fractures, typically measured in megapascals (). For engineering thermoplastics like (polyamides), tensile strength commonly ranges from 50 to 90 , enabling their use in load-bearing parts such as gears and cables. In contrast, commodity polymers like (HDPE) exhibit lower values around 15-30 , reflecting their role in less demanding applications like packaging. Young's modulus, also known as the , quantifies a polymer's in the linear and is defined as the of to : E = \frac{\sigma}{\epsilon} where \sigma is and \epsilon is . For semi-crystalline polymers, such as polyamides or polyethylenes, Young's modulus typically falls between 1 and 3 GPa, indicating moderate due to the reinforcing effect of crystalline domains. This value is determined through and highlights how chain alignment and crystallinity enhance resistance to deformation without permanent damage./04:_Mechanical_Properties) Elongation at break measures the of a polymer, expressed as the increase in length from the original before . Elastomers, such as polyurethanes or , often achieve elongations exceeding 500%, allowing extreme stretching and recovery, which is essential for , tires, and biomedical devices. This high extensibility arises from flexible, cross-linked molecular networks that store and release efficiently. Viscoelasticity in polymers refers to their combined viscous and elastic responses, leading to time-dependent deformation under load. is the gradual increase in over time under constant , while is the decrease in under fixed ; both phenomena are prominent in amorphous and semi-crystalline polymers at . These behaviors, modeled by elements like or Kelvin-Voigt in rheological analysis, influence long-term performance in applications like adhesives and composites, where sustained loads can lead to dimensional changes. , such as the degree of crystallinity, can modulate viscoelastic effects by altering chain mobility.

Thermal and Phase Behavior

Polymers exhibit distinct thermal transitions that govern their phase behavior, primarily influenced by their molecular structure and composition. The (Tg) marks the reversible shift in amorphous polymers from a rigid, glassy state to a flexible, rubbery state, where segmental mobility increases significantly. For instance, displays a Tg of approximately 100°C, allowing it to maintain rigidity at while softening upon heating. This transition is critical for applications requiring dimensional stability below Tg and elasticity above it. In copolymers, can be predicted using the Fox equation, which accounts for the weight fractions of individual components:
\frac{1}{T_g} = \sum \frac{w_i}{T_{g i}}
where w_i is the weight fraction of the ith component and T_{g i} is its temperature. This empirical relation assumes additive contributions from homopolymer segments, providing a useful for random copolymers without strong interactions. Experimental validations confirm its applicability in many systems, though deviations occur in block copolymers due to .
Crystalline polymers undergo at the melting temperature (Tm), where ordered lamellae disrupt into a disordered melt, distinct from the Tg of amorphous regions. (HDPE), for example, melts around 130°C, enabling processing via or molding while preserving mechanical integrity below this point. Crystallization from the melt requires , where the temperature drops below Tm to drive and growth; greater accelerates kinetics but can yield metastable structures with reduced perfection, influencing final and properties. The of polymer blends is described by Flory-Huggins theory, which models the of mixing through the interaction parameter χ, quantifying enthalpic differences between unlike segment contacts. Miscibility occurs when χ is sufficiently low (typically χ < 0.5 for symmetric blends at equilibrium), promoting a single phase; higher values lead to , as seen in immiscible polystyrene-poly() blends. This parameter, often temperature-dependent, guides blend design for tailored thermal behavior. Plasticizers enhance flexibility by reducing intermolecular forces, thereby lowering and enabling use in rigid polymers like (PVC). Phthalate esters, such as di(2-ethylhexyl) phthalate, are commonly added to PVC at 30-50 wt% to depress from ~80°C to below 0°C, transforming it into a pliable for cables and . This effect arises from the plasticizer's of polymer chains, increasing free volume without altering primary structure.

Electrical, Optical, and Chemical Properties

Polymers exhibit a range of electrical properties depending on their molecular structure and composition. The dielectric constant, a measure of a material's ability to store in an , typically ranges from 2 to 4 for common non-polar polymers such as (2.1) and (2.3), while polar polymers like can reach up to 3.5. Conjugated polymers, such as , demonstrate electrical conductivity through delocalized π-electrons along their backbone, with doped forms achieving conductivities of 0.1 to 4 S/cm under optimal conditions like acidic doping and specific synthesis methods. (PVDF) displays , generating an electric charge in response to mechanical stress due to its polar β-phase crystalline structure, making it suitable for sensors and actuators. Optical properties of polymers are influenced by their chain packing and electronic structure. Poly(methyl methacrylate) (PMMA) offers high with approximately 92% in the visible range (380-780 nm), resembling and enabling applications in optical lenses. Oriented polymer chains induce , where the refractive indices differ along and perpendicular to the chain direction due to anisotropic , often quantified as Δn ≈ 0.1-0.2 for stretched films. Most polymers have a around 1.5, as seen in (1.59) and PMMA (1.49), which governs bending and reflection in optical components. Chemical properties determine polymer stability and interactions with environments. Resistance to solvents is predicted using solubility parameters, which separate cohesive energy into dispersion (δ_d), polar (δ_p), and hydrogen-bonding (δ_h) components; for instance, has δ_d ≈ 18.6 MPa^{1/2}, δ_p ≈ 1.0 MPa^{1/2}, and δ_h ≈ 2.0 MPa^{1/2}, indicating solubility in non-polar solvents like . Polyesters, such as , undergo where ester linkages react with water under acidic or basic conditions to form carboxylic acids and alcohols, a process accelerated by heat or enzymes and leading to chain scission. These properties have enabled practical advancements, including the commercialization of organic light-emitting diodes (OLEDs) using conductive polymers like poly(3,4-ethylenedioxythiophene) (PEDOT) post-2000, where their and facilitate efficient charge injection and light emission in flexible displays.

Applications

Industrial and Consumer Applications

Polymers are integral to industrial and consumer applications, leveraging their versatile properties for durability, lightweighting, and cost-effectiveness across sectors. In , (PE) films dominate due to their excellent moisture barrier and flexibility, enabling efficient protection and preservation of goods. Global PE totaled approximately 108 million metric tons in 2023, with over half allocated to uses such as flexible films and wraps. (PET) complements this by providing clarity and strength for rigid containers, particularly bottles for beverages and consumer products. Worldwide PET bottle reached about 28 million metric tons in 2024, underscoring its prevalence in single-use and reusable formats. Construction relies on polymers for structural integrity and , with (PVC) pipes serving as a cornerstone for , , and systems owing to their resistance and longevity. Global PVC pipe was 25.9 million metric tons in 2024, reflecting sustained demand in residential and infrastructure projects. (PU) foams further enhance building performance through superior , reducing in heating and cooling. Total global PU exceeded 21 million metric tons in 2023, with rigid PU foams—primarily used for —accounting for a substantial share valued at USD 22.76 billion in 2024. In the automotive industry, polymers contribute to safety, performance, and sustainability, notably through in , which provides essential traction and resistance. The tire sector consumes roughly 70% of global rubber output, including a significant portion of production, valued at USD 24.29 billion in 2024 and estimated at around 15 million metric tons annually. Polymer composites, including reinforced thermoplastics, enable lightweighting of components like bumpers and panels, improving fuel economy and range. The automotive polymer composites market was valued at USD 10.20 billion in 2023, driven by regulatory pressures for emissions reduction. Consumer goods benefit from polymers' adaptability in everyday items, with fibers leading in textiles for apparel, , and home furnishings due to their wrinkle resistance and ease of care. comprised 57% of global , which hit 132 million metric tons in 2024, translating to roughly 75 million metric tons of output. Polymer adhesives, often based on acrylics or urethanes, bond diverse materials in products like , , and , enhancing assembly efficiency. The global adhesives and sealants market, heavily reliant on polymers, reached USD 82.88 billion in 2024. The polymer sector's scale is evident in its resource intensity, accounting for about 8% of global use as feedstock in 2024, primarily for and derivatives. Looking to 2025, projections emphasize transitions to models, with investments in recycled polymer technologies anticipated to surge toward USD 100 billion by 2030 to address waste and support sustainable supply chains. In 2025, the adoption of bio-based and recycled polymers in and automotive sectors has accelerated, driven by new regulations on waste.

Biomedical and Biological Applications

Polymers play a pivotal role in biomedical and biological applications due to their , tunable degradation profiles, and ability to mimic biological tissues. These materials enable innovations in , tissue regeneration, and implantable devices, improving therapeutic outcomes while minimizing immune responses. Synthetic polymers like (PEG) and poly(lactic-co-glycolic acid) () are particularly valued for their versatility in creating structures that interact safely with physiological environments. In , hydrogels based on have emerged as effective carriers for controlled release, leveraging their high water content and swelling properties to encapsulate and gradually release therapeutic agents. hydrogels facilitate sustained of drugs such as chemotherapeutics, reducing dosing frequency and systemic in cancer therapy. For instance, injectable -based systems can form at tumor sites, achieving prolonged release over days to weeks while maintaining . Recent formulations incorporate stimuli-responsive elements, like pH-sensitive linkages, to trigger release in acidic tumor microenvironments, enhancing precision. Tissue engineering relies on biodegradable scaffolds from polymers like polylactic acid (PLA) and PLGA to support cell growth and tissue regeneration. These scaffolds provide mechanical support and degrade into non-toxic byproducts, allowing gradual replacement by native in applications such as and repair. PLA scaffolds, often modified with bioactive molecules, promote adhesion and proliferation, with degradation rates tunable from months to years via copolymer ratios. PLGA variants offer faster , making them suitable for soft tissue engineering, where they enhance vascularization and deposition. For implants, silicone elastomers are widely used in prosthetics and long-term devices due to their flexibility, durability, and low toxicity. These materials form soft, biocompatible components in breast implants and maxillofacial prosthetics, withstanding mechanical stress while resisting degradation in vivo. Silicone's inert nature minimizes inflammation, enabling safe integration over decades. Hydrogels, including silicone-infused variants, are staples in contact lenses, providing oxygen permeability and moisture retention to prevent corneal hypoxia during extended wear. Modern silicone hydrogel lenses achieve high water content (up to 80%) and modulus values below 1 MPa, improving comfort and reducing infection risks. Recent advances highlight polymers in nucleic acid delivery, such as coatings on nanoparticles (LNPs) for mRNA vaccines. In the , PEG-lipid conjugates stabilized LNPs in vaccines, extending circulation time and boosting immunogenicity by evading immune clearance. These coatings reduced lipid content while maintaining efficacy, with formulations achieving over 90% in dendritic cells. For , polymer-based systems like lipid-polymer hybrids and modified chitosans have improved delivery efficiency, enabling targeted edits with minimal off-target effects. By 2025, these carriers demonstrated up to 70% editing for therapeutic applications like genetic disorders, addressing limitations in viral vectors through biodegradability and scalability.

Characterization

Molecular and Structural Characterization

Molecular and structural characterization of polymers involves a suite of analytical techniques designed to elucidate their , chain length distribution, , branching, and overall architecture, which are critical for understanding their behavior and performance. These methods provide insights into the molecular-level features that define a polymer's , distinct from macroscopic properties. Key approaches include spectroscopic, chromatographic, microscopic, and techniques, each offering complementary information on structure at different scales. Nuclear magnetic resonance (NMR) is a cornerstone for determining polymer microstructure, particularly and monomer sequence distribution. In 13C NMR, the chemical shifts of carbon atoms in the polymer backbone reveal the stereochemical configuration; for instance, in , the relative intensities of methyl carbon resonances distinguish isotactic, syndiotactic, and atactic forms, with isotactic showing a single sharp peak at around 21.8 ppm due to its regular stereoregularity. This technique, pioneered in early studies on stereoregular polymers, allows quantitative assessment of by integrating peak areas, achieving resolutions down to dyad or levels for copolymers. Proton () NMR complements this by identifying end groups and sequence defects in copolymers like ethylene-propylene. For complex architectures such as branched or star polymers, multidimensional NMR variants like DOSY (diffusion-ordered ) map chain dimensions and connectivity. Infrared (IR) spectroscopy, including Fourier-transform IR (FTIR), excels at identifying functional groups and overall composition in polymers, providing a rapid, non-destructive means to confirm chemical identity. Characteristic absorption bands, such as the C=O stretch at 1700-1750 cm⁻¹ for polyesters or the C-H stretch at 2800-3000 cm⁻¹ for polyolefins, enable qualitative and quantitative analysis of additives, cross-links, or degradation products. In copolymers, IR distinguishes monomer ratios by band intensities, as seen in ethylene-vinyl acetate where the acetate carbonyl peak quantifies vinyl acetate content up to 40 mol%. (ATR) FTIR extends this to solid samples, facilitating in-situ monitoring of structural changes during processing. Gel permeation chromatography (GPC), also known as (SEC), is the primary method for characterizing molecular weight distribution (MWD) and polydispersity index (PDI), essential for assessing chain length heterogeneity. Polymers are separated by hydrodynamic volume in a solvent-eluting column packed with porous beads; larger chains elute first, and calibration with standards (e.g., ) yields number-average (M_n) and weight-average (M_w) molecular weights via the Mark-Houwink relation, [η] = K M^a, where correlates to size. For example, in , GPC reveals bimodal distributions from blending, with PDI values above 2 indicating broad MWD that impacts melt flow. Advanced multi-detector GPC couples light scattering and viscometry for absolute without standards, particularly useful for branched polymers like . Atomic force microscopy (AFM) provides nanoscale visualization of individual polymer chains and their conformations, bridging molecular and morphological scales. In tapping mode, AFM images single-chain adsorption on substrates, revealing coil or extended conformations; for on , chain radii of gyration (R_g) measured around 10-20 nm align with theoretical models. This technique probes in ultrathin films or solutions, detecting branching via irregular chain outlines, and has been instrumental in confirming bottlebrush polymer architectures with side-chain grafting densities. Transmission electron microscopy (TEM) offers high-resolution imaging of polymer at the nanoscale, particularly for block copolymers and nanocomposites. Cryo-TEM preserves native structures in vitrified samples, visualizing lamellar or cylindrical microdomains with periodicities of 10-100 nm, as in polystyrene-block-polybutadiene where is evident from contrasting electron densities. Staining with enhances visibility of unsaturated segments, enabling 3D reconstructions via for complex architectures like micelles. Small-angle X-ray scattering (SAXS) and (WAXS) probe crystallinity and larger-scale structures non-destructively. SAXS detects long-range order, such as lamellar stacking in semicrystalline polymers, yielding d-spacings from , λ = 2d sinθ; in , SAXS long periods of 20-30 nm indicate crystal-amorphous alternation. WAXS resolves atomic-scale crystallinity through peaks, quantifying percent crystallinity (e.g., 50-70% for ) via peak . Combined SAXS/WAXS analysis, as in studies of poly(ε-caprolactone), reveals how processing affects perfection and orientation. These techniques reference microstructure elements like size without delving into their formation mechanisms.

Property Characterization

Property characterization of polymers involves a suite of standardized techniques to quantify their mechanical, thermal, rheological, and electrical behaviors, enabling precise material specification and performance prediction. These methods focus on functional responses under controlled conditions, distinct from structural analyses that probe molecular composition. Mechanical properties, such as tensile modulus and strength, are commonly assessed through tensile testing, where a polymer specimen is subjected to uniaxial tension until failure. The ASTM D638-22 standard outlines this procedure for unreinforced and reinforced plastics, using dumbbell-shaped samples to determine key metrics like Young's modulus, yield strength, and elongation at break, which reflect the material's stiffness and ductility under load. For example, in polyethylene testing, modulus values typically range from 200 to 1000 MPa, establishing baseline rigidity for packaging applications. Thermal properties are evaluated using (DSC) to identify temperature (Tg) and melting temperature (Tm), which govern phase changes and processability. In DSC, a polymer sample is heated at a controlled rate while measuring flow differences relative to a reference, revealing endothermic or exothermic transitions; Tg appears as a step change in , often between -100°C and 200°C for common polymers like (Tg ≈ 100°C). (TGA) complements this by monitoring mass loss with temperature to determine onset, typically above 300°C for stable thermoplastics, aiding in assessing thermal stability and filler content. For instance, TGA of ,6 shows degradation starting at 482°C with near-complete mass loss. Rheological characterization, particularly melt viscosity, is performed via viscometry to measure the (MFI), indicating processability in or molding. Under ASTM D1238, a molten polymer is extruded through a under specified and load, with MFI reported in grams per 10 minutes; low-MFI materials like (≈0.2 g/10 min) exhibit higher suitable for . This single-point measurement provides a practical proxy for molecular weight and flow behavior without full shear-rate profiling. Electrical properties, including , are probed using , which applies an across frequencies to model . This technique generates Nyquist plots to separate bulk resistance from interfacial effects, yielding values often in the 10^{-12} to 10^{-6} S/cm range for insulating polymers, with enhancements via conductive fillers. In , it reveals ionic contributions dominant at low frequencies, crucial for applications.

Degradation and Environmental Impact

Types of Degradation

Polymers undergo degradation through various mechanisms that compromise their structural integrity and performance, primarily at the molecular level via chain scission, cross-linking, or . These processes are influenced by environmental factors such as , , , and oxygen, leading to changes in mechanical properties like reduced tensile strength and increased . The main types include , , hydrolytic and oxidative degradation, as well as associated product failure modes. Thermal degradation occurs when polymers are exposed to elevated temperatures, typically above their (Tm), resulting in breakage and volatile product formation. In this process, chain scission predominates, where covalent bonds along the polymer backbone cleave, reducing molecular weight and causing material weakening. For instance, in (PMMA), degradation above 220°C initiates random unzipping , primarily yielding units through end-chain and propagation. This mechanism contrasts with random scission seen in at temperatures exceeding 450°C, which produces a of hydrocarbons without predominant recovery. Overall, varies with polymer structure; aromatic polymers like exhibit higher resistance due to stabilized bonds, while aliphatic chains degrade more readily. Photodegradation is triggered by (UV) radiation, particularly in the 290–400 nm range, which excites chromophoric groups in the polymer, generating free radicals that propagate chain reactions. In polyolefins such as , UV absorption leads to formation and subsequent radical-mediated scission, causing surface erosion, yellowing, and embrittlement over time. The process often involves photo-oxidative pathways, where oxygen reacts with radicals to form carbonyl groups, further reducing ductility and increasing fragility, as observed in (HDPE) films after prolonged exposure. This degradation is more pronounced in unstabilized polymers, with depth limited to surface layers unless antioxidants are absent. Hydrolytic degradation involves the cleavage of susceptible bonds, such as in polyesters, by molecules, often catalyzed by acids or bases, leading to chain shortening and loss of mechanical integrity. In aliphatic polyesters like poly() (), proceeds via nucleophilic attack, producing and end groups that autocatalyze further breakdown, resulting in bulk erosion and reduced toughness. This mechanism is pH-dependent and accelerated in humid environments, with poly(ethylene terephthalate) (PET) showing slower rates due to its semi-crystalline structure. Oxidative degradation, including auto-oxidation, arises from reactions with atmospheric oxygen, forming peroxyl radicals that initiate chain propagation and termination. In rubbers like or rubber, auto-oxidation begins with hydrogen abstraction at allylic positions, leading to double-bond scission, cross-linking, and hardening, which diminishes elasticity. The process exhibits an induction period before rapid , influenced by temperature and trace metals that catalyze formation. Product failure modes often stem from these degradative mechanisms under applied . Creep rupture involves time-dependent deformation under constant load below the strength, culminating in due to viscoelastic and chain disentanglement, as seen in PMMA components under prolonged tension. Environmental cracking (ESC) in HDPE pipes occurs when tensile combines with chemical agents like , inducing craze formation and crack propagation through localized chain slippage and void growth. These modes highlight how accelerates failure in load-bearing applications.

Environmental and Sustainability Considerations

Polymers, particularly synthetic ones like and (), contribute significantly to environmental pollution through the release of , which arise from the fragmentation of larger debris via processes. An estimated 14 million metric tons of , much of it in microplastic form, enters the oceans annually as of 2025, posing risks to ecosystems and entering the global . plastics exhibit high environmental persistence, remaining intact in environments for decades to hundreds of years due to their resistance to natural breakdown under ambient conditions. While most conventional polymers are non-biodegradable in natural settings, certain bio-based alternatives like () can degrade under specific controlled conditions. biodegrades completely in industrial composting environments, where temperatures of 58–60°C and high humidity facilitate microbial activity, achieving up to 90% mineralization to CO₂ within 70–180 days. However, does not readily break down in home composting, , or settings without such optimized conditions, limiting its environmental benefits if not properly managed. Recycling remains a core strategy for mitigating polymer waste, encompassing mechanical methods like re-extrusion of sorted plastics into new products and chemical approaches such as to recover monomers for repolymerization. These processes support objectives, with the targeting 55% of plastic packaging waste by 2030 to reduce landfill and incineration reliance. Enzymatic has seen notable 2025 advances, including engineered enzymes that achieve up to 65% energy savings and cost reductions in PET , enabling more efficient breakdown at milder conditions compared to traditional chemical methods. Sustainable polymer alternatives emphasize bio-based feedstocks to lessen dependency and emissions. For instance, bio-based (bio-PE) derived from offers identical properties to conventional PE while capturing CO₂ during plant growth, reducing lifecycle by up to 70%. These innovations, alongside ongoing developments, address gaps in traditional by promoting scalable, low-impact pathways for polymer production and end-of-life management. Internationally, negotiations for a global plastics under the UN Programme's Intergovernmental Negotiating (INC) continued into 2025, aiming for a legally to end , though the fifth session (INC-5) adjourned without agreement in August 2025 and was set to reconvene later that year.

Nomenclature

Standardized Nomenclature

The International Union of Pure and Applied Chemistry (IUPAC) establishes standardized nomenclature for polymers to ensure unambiguous description of their chemical structures, prioritizing structure-based naming for precision while allowing source-based naming for simplicity when the monomer is clearly identifiable. Source-based nomenclature derives the polymer name directly from the monomer or monomers used in synthesis, prefixing "poly" to the monomer name enclosed in parentheses, such as poly(ethylene) for the polymer formed from ethylene monomers. In contrast, structure-based nomenclature employs the constitutional repeating unit (CRU)—the smallest structural motif that repeats to form the polymer chain—and names the polymer as poly followed by the CRU in square brackets, for example, poly(oxyethylene) for polyethylene oxide, which highlights the repeating -O-CH₂-CH₂- unit rather than the source material. For copolymers, IUPAC recommends connective prefixes to denote the arrangement of repeating units, such as "co" for random or unspecified copolymers (e.g., poly(styrene-co-butadiene)), "" for block copolymers (e.g., poly(styrene--butadiene)), and "" for alternating copolymers (e.g., poly(styrene--maleic anhydride)). These connectives are italicized and placed between the names of the components, which are listed in alphabetical order, enabling clear differentiation of architectural features without implying specific sequencing unless further specified. To address regularity, including , IUPAC employs "source-with-regularity" nomenclature for polymers with ordered , incorporating stereodescriptors into the name to specify configurations along the chain; for instance, isotactic is denoted as it-poly(propene), using the 'it-' prefix to indicate isotactic configuration. This approach extends structure-based rules for irregular polymers by using slashes to separate multiple CRUs (e.g., poly(but-1-ene-1,4-diyl/1-vinylethane-1,2-diyl) for polymers with irregular constitutional units), while tactic polymers like syndiotactic ones use descriptors such as "rac" or "meso" for specifications. The 2017 IUPAC recommendations extend these principles to complex architectures, such as and dendritic polymers, introducing substitutive and multiplicative naming systems based on units, dendrons, and generations; for example, polymers are denoted following general rules as star-poly(A), while dendritic structures use detailed CRU assemblies like α,α′,α″-[ethane-1,1,1-triyltri(4,1-phenylene)]tris[ω-hexadecahydro-dendro G4-(oxymethylenebenzene-1,3,5-triyl)] for precise generational control. Hyperbranched polymers follow substitutive rules with "hyper-" prefixes, naming them as α-(core)-ω-(end-group)-hyper-poly(CRU), such as α-(propane-1,1,1-triyl)-ω-(hydroxymethyl)-hyper-poly[methyleneoxy(2-methyl-1-oxoethane-1,2,2-triyl)].

Common Naming Conventions

Common naming conventions for polymers often rely on source-based , where the name is derived from the or monomers used in , prefixed with "poly" to indicate the polymeric nature. For instance, (PE) is named after its ethylene , while (PVC) reflects the precursor. These generic names are widely adopted in and for their simplicity and direct connection to . Trade names, assigned by manufacturers, provide branded identifiers for commercial polymers and are frequently used in everyday contexts, marketing, and product specifications. Examples include for polytetrafluoroethylene (PTFE), known for its non-stick properties; for a type of fiber valued in protective gear; and [Lucite](/page/polymethyl methacrylate) for polymethyl methacrylate (PMMA), a clear material. Other notable trade names are Dacron and Mylar, both referring to poly(ethylene terephthalate) (PET) in textile and film applications, respectively. These names enhance market recognition but can lead to confusion without chemical equivalents. Abbreviations serve as shorthand in technical writing, research papers, and patents to streamline communication. Standard examples include PS for , PET for poly(ethylene terephthalate), and PMMA for polymethyl methacrylate, following guidelines that prioritize brevity while maintaining clarity. The International Union of Pure and Applied Chemistry (IUPAC) endorses such abbreviations, recommending their full expansion on first use to avoid ambiguity. Historically, some trade names have evolved into generic terms due to widespread adoption, illustrating shifts in nomenclature over time. , trademarked in 1907 by inventor for the first fully synthetic plastic (a phenol-formaldehyde resin), became synonymous with phenolic resins in common parlance despite its proprietary origins. This phenomenon, similar to "" for polyamides, highlights how influential polymers can transcend branding to influence everyday language.

References

  1. [1]
    What are polymers? - IUPAC | International Union of Pure and ...
    Polymers are substances composed of macromolecules, very large molecules with molecular weights ranging from a few thousand to as high as millions of grams/mole ...Missing: sources | Show results with:sources
  2. [2]
    Polymer Fundamentals - Chemistry LibreTexts
    Jan 22, 2023 · Polymers are long chain, giant organic molecules are assembled from many smaller molecules called monomers. Polymers consist of many repeating ...Missing: authoritative sources
  3. [3]
    Polymers - MSU chemistry
    Polymers formed by a straightforward linking together of monomer units, with no loss or gain of material, are called addition polymers or chain-growth polymers.3. Properties Of... · Synthesis Of Addition... · Condensation PolymersMissing: authoritative | Show results with:authoritative
  4. [4]
    Polymers in our daily life - PMC - NIH
    Polymers, a large class of materials, consist of many small molecules named monomers that are linked together to form long chains and are used in a lot of ...Missing: authoritative | Show results with:authoritative
  5. [5]
    The Origin of the Polymer Concept
    No readable text found in the HTML.<|control11|><|separator|>
  6. [6]
    From Polymer to Macromolecule: Origins and Historical Evolution of ...
    May 21, 2025 · Polymer science is one such subset of chemistry with its own dialect, with the core concept of the polymer dating back to 1832. ... The current ...
  7. [7]
    PLASTOMER Definition & Meaning - Merriam-Webster
    a relatively tough usually hard and rigid polymeric substance compare elastomer, plastic entry 2 sense 3. Word History. Etymology. plasto- + -mer ...
  8. [8]
    When Did People Start Making Silk? - History.com
    Sep 24, 2025 · Chinese legend says that silk was discovered in about 2700 B.C. by the Empress Leizu, wife of the mythical Yellow Emperor, when she sipped tea ...
  9. [9]
    Rubber processed in ancient Mesoamerica, MIT researchers find
    Jul 14, 1999 · An MIT research team that includes an undergraduate has demonstrated that by 1600 BC, an ancient civilization was processing latex to produce rubber.
  10. [10]
    Charles Goodyear and the Vulcanization of Rubber
    Charles Goodyear's discovery of the vulcanization of rubber—a process that allows rubber to withstand heat and cold—revolutionized the rubber industry in the ...
  11. [11]
    The Age of Plastic: From Parkesine to pollution | Science Museum
    Oct 11, 2019 · Its inventor, the Birmingham-born artisan-cum-chemist Alexander Parkes, patented this new material in 1862 as Parkesine. Considered the first ...
  12. [12]
    Hermann Staudinger Foundation of Polymer Science - Landmark
    This new concept, referred to as "macromolecules" by Staudinger in 1922, covered both synthetic and natural polymers and was the key to a wide range of modern ...
  13. [13]
    Wallace Carothers and the Development of Nylon - Landmark
    One of these, synthesized from adipic acid and hexamethylenediamine on February 28, 1935, was called fiber 66 because each of its components had six carbon ...
  14. [14]
    The Nobel Prize in Chemistry 1963 - NobelPrize.org
    The Nobel Prize in Chemistry 1963 was awarded jointly to Karl Ziegler and Giulio Natta "for their discoveries in the field of the chemistry and technology ...Missing: 1950s | Show results with:1950s
  15. [15]
    The Nobel Prize in Chemistry 2000 - NobelPrize.org
    The Nobel Prize in Chemistry 2000 was awarded jointly to Alan J. Heeger, Alan G. MacDiarmid and Hideki Shirakawa for the discovery and development of ...
  16. [16]
    The Rise of Polyhydroxyalkanoates (PHA) in Packaging: A Unique ...
    Mar 26, 2025 · Unlike fossil-based plastics, PHA has a low carbon footprint and is both biodegradable and compostable in all environments, making it an ideal ...
  17. [17]
    Braskem unveils bio-based product innovations at K 2025
    Oct 9, 2025 · Braskem's collaboration with Dutch innovators Bottle Up and Eurobottle brings bio-based materials into practical, reusable designs.
  18. [18]
    Biopolymer: A Sustainable Material for Food and Medical Applications
    Plants, animals, microorganisms, and agricultural wastes are examples of natural biological sources of biopolymers. Plant sources, such as rice, maize [8], ...
  19. [19]
    Natural Polymer - an overview | ScienceDirect Topics
    Natural polymers are polymers found in nature, extracted from plants and materials. Examples include starch, wool, proteins, DNA, cellulose, and chitin.
  20. [20]
    Polymers | Research Starters - EBSCO
    In nature, polymers play critical roles in processes such as information storage, catalysis of biochemical reactions, structural support, and energy storage.Overview · Polymers In Nature · Context
  21. [21]
    Cellulose - American Chemical Society
    Jan 19, 2009 · Cellulose is the most abundant organic compound on Earth. It is the main constituent of plant fiber; plants contain on average 33% cellulose, and cotton is the ...Missing: biomass percentage
  22. [22]
    A Review on the Modification of Cellulose and Its Applications - MDPI
    1.2.​​ All plant matter has a cellulose concentration of roughly 33%, on average. It is possible to extract cellulose from its raw biomass materials, and it has ...
  23. [23]
    A Review of Natural Polysaccharides: Sources, Characteristics ...
    Natural polysaccharides, which are described in this study, are some of the most extensively used biopolymers in food, pharmaceutical, and medical applications.
  24. [24]
    Anatomy, Connective Tissue - StatPearls - NCBI Bookshelf - NIH
    Mar 5, 2023 · [8] Tendons are connective tissue structures comprised of a hierarchical arrangement of collagen molecules that arrange into collagen ...
  25. [25]
    Biological Molecules | Biology I - Lumen Learning
    There are four major classes of biological macromolecules (polymers), and each is an important component of the cell and performs a wide array of functions.
  26. [26]
    Lipid Composition of Latex and Rubber Particles in Hevea ... - MDPI
    Natural rubber is usually synthesized in the rubber particles present in the latex of rubber-producing plants such as the Pará rubber tree (Hevea ...
  27. [27]
    [PDF] Lignins: Structure and distribution in wood and pulp
    Lignin is the stuff that makes trees “woody.” Usually constituting from one-fifth to one-third of wood, lignin strongly influences its chemical and physical ...
  28. [28]
    Lignocellulosic biomass and its main structural polymers as ...
    Jun 12, 2025 · Lignocellulosic biomass typically comprises approximately 90% organic macromolecules (cellulose, hemicellulose, and lignin) which form the ...
  29. [29]
    2.3 Biological Molecules – Concepts of Biology – 1st Canadian Edition
    There are four major classes of biological macromolecules (carbohydrates, lipids, proteins, and nucleic acids), and each is an important component of the cell.
  30. [30]
    Collagen: What It Is, Types, Function & Benefits - Cleveland Clinic
    Collagen accounts for 30% of your body's protein. It provides structure, support or strength to your skin, muscles, bones and connective tissues.
  31. [31]
    Micro-organisms in latex and natural rubber coagula of Hevea ...
    Natural rubber, produced by coagulation of the latex from the tree Hevea brasiliensis, is an important biopolymer used in many applications for its ...
  32. [32]
    Synthetic Polymer - an overview | ScienceDirect Topics
    1.4.​​ Synthetic polymers are defined as polymers that are artificially produced in laboratories. These are also known as man-made polymers.
  33. [33]
    Petroleum-Based Polymer - an overview | ScienceDirect Topics
    Petroleum-based polymers are defined as synthetic polymers derived from petroleum, which can possess properties similar to other materials and are utilized in ...
  34. [34]
    Synthetic Organic Polymers – Introductory Chemistry
    Synthetic polymers are human-made polymers. They can be classified into four main categories: thermoplastics, thermosets, elastomers, and synthetic fibers.Missing: major | Show results with:major
  35. [35]
    Polyethylene (PE Plastic) – Structure, Properties & Toxicity
    Jul 9, 2025 · Polyethylene (PE) is one of the most popular thermoplastic materials. It is available in different crystalline structures, referred to as HDPE, LDPE, and LLDPE.
  36. [36]
    Thermoset Epoxies | CompositesWorld
    Epoxy is a thermosetting matrix resin and among the most commonly used resin systems in the composites industry.
  37. [37]
    What is SBR (Styrene-Butadiene) Rubber?
    SBR (Styrene butadiene) rubber is one of the first synthetic rubber materials ever invented. It has many of the same properties as natural rubber.
  38. [38]
    Polystyrene - Plastics Europe
    Polystyrene is a synthetic polymer made from styrene monomer, which is a liquid petrochemical. It is a thermoplastic polymer that softens when heated.
  39. [39]
    What is polyurethane?
    Polyurethanes are plastic polymers made by combining diisocyanates ( TDI and MDI) and polyols. There are literally hundreds of different types of polyurethanes ...
  40. [40]
    4D printing of shape memory polylactic acid (PLA) - ScienceDirect.com
    Sep 16, 2021 · Fig. 2 displays the place of PLA between other biodegradable polymers. PLA was discovered in 1845 but commercialized in the early 1990s by the ...
  41. [41]
    Biodegradable Polylactic Acid and Its Composites - NIH
    Jul 19, 2023 · Polylactic acid (PLA) is a biodegradable bio-based aliphatic polyester that can be extracted from 100% renewable resources, such as corn ...
  42. [42]
    Synthetic Polymers | Organic Chemistry Class Notes - Fiveable
    Their properties depend on chemical composition, molecular weight, and structure, allowing for customization to meet specific needs. Key types include ...
  43. [43]
    [PDF] A Brief Guide to Polymer Nomenclature | IUPAC
    It is defined as the ratio of the mass-average molar mass (Mm) to the number-average molar mass (Mn) i.e. Ð = Mm/Mn.4 Symbols for physical quantities or ...
  44. [44]
    Introduction to proteins and amino acids (article) - Khan Academy
    Each bond forms in a dehydration synthesis (condensation) reaction. During ... proteins are polymers of amino acids whereas amino acid is a monomer of protein.
  45. [45]
    [PDF] A Brief Guide to Polymerization Terminology (IUPAC Technical ...
    In ring-opening polymerization (ROP), the repeating units contain fewer rings than the monomer. In cyclopolymerization, the repeating units contain more rings ...
  46. [46]
    Polymer Architecture - an overview | ScienceDirect Topics
    Every natural, seminatural, and synthetic polymer falls into one category of architecture: linear, graft, branched, cross-linked, block, star-shaped, or dendron ...
  47. [47]
    The Beauty of Branching in Polymer Science | Macromolecules
    May 12, 2020 · Branching can range from hard to detect side reactions during the growth of linear polymers to perfectly branched, fractal-like dendrimers. As ...Missing: microstructure | Show results with:microstructure
  48. [48]
    Polymer Architecture - Leonard Gelfand Center - Carnegie Mellon ...
    This structure is called a branched polymer chain. The following figure shows the case of low density polyethylene. Notice the ethylene segments dangling ...
  49. [49]
  50. [50]
    Molecular weight - DoITPoMS
    The polydispersity index is defined as the ratio of the weight average molecular weight to the number average molecular weight, and it gives a measure of the ...Missing: PDI | Show results with:PDI
  51. [51]
    31.3 Copolymers - Organic Chemistry | OpenStax
    Sep 20, 2023 · Block copolymers are those in which different blocks of identical monomer units alternate with each other; graft copolymers are those in which ...
  52. [52]
    Alternating Copolymer - an overview | ScienceDirect Topics
    Copolymers can be distinguished into alternating-, random-, graft-, and block copolymers. Alternating copolymers may be considered as homopolymers with a ...
  53. [53]
    24.4: Ziegler–Natta Catalysts and Polymer Stereochemistry
    Jun 5, 2019 · Ziegler-Natta catalysts are prepared by reacting certain transition metal halides with organometallic reagents such as alkyl aluminum, lithium and zinc ...
  54. [54]
    The Influence of Ziegler-Natta and Metallocene Catalysts on ...
    Ziegler and Natta are both awarded the Nobel Prize for Chemistry 1963 in recognition of their work on the Ziegler-Natta catalyst. 1957, Commercial production of ...
  55. [55]
    The Configuration of Real Polymer Chains - AIP Publishing
    Polymer configuration is influenced by spatial interferences, and the power of chain length is greater than 0.50, approaching 0.60 for long chains in good ...Missing: conformation seminal
  56. [56]
    [PDF] Lecture 16 Morphology of Semicrystalline Polymers February 23, 2001
    Feb 23, 2001 · Random coil conformations occur in solution and the melt. • Ordered chain conformations exist in crystals. • Extended chain conformations ...
  57. [57]
    Crystallinity / Degree of Crystallinity - NETZSCH Analyzing & Testing
    The melting enthalpy for 100% crystalline PE is 293 J/g and is 207 J/g for PP. This yields a calculated crystallinity for LDPE of 46.5%, for HDPE of 74.2% and ...
  58. [58]
    The measurement of the crystallinity of polymers by DSC
    The degree of crystallinity is then defined as X c = Δ H f (T m )/ Δ H f o T m o where Xc is the weight fraction extent of crystallinity, ΔHf(Tm) is the ...
  59. [59]
    A Phenomenological Theory of Spherulitic Crystallization
    To account for spherulitic crystallization from the melt, one must explain the origins (i) of fibrous crystal habits in the absence of appreciable temperature ...
  60. [60]
  61. [61]
    Crystallization of Poly(ethylene terephthalate): A Review - PMC - NIH
    In fact, it can be easily attained in an amorphous form by quenching from its molten state [22,23,24,25], since a cooling rate of 1 K/s is sufficient to ...
  62. [62]
    Toward Living Radical Polymerization - ACS Publications
    Radical polymerization is one of the most widely used processes for the commercial production of high-molecular-weight polymers.
  63. [63]
    [PDF] Radical Polymerization: Kinetics and Mechanism
    ... free-radical polymerization and of controlled radical polymerization. The first SML meeting was organized by Ken O'Driscoll and Saverio. Russo at Santa ...
  64. [64]
    Living Anionic Polymerization - an overview | ScienceDirect Topics
    Living anionic polymerization is defined as a type of polymerization that occurs in the absence of chain transfer and chain termination, allowing for ...Missing: cationic | Show results with:cationic
  65. [65]
    Living Anionic Polymerization Celebrates 60 Years: Unique ...
    Jun 19, 2017 · Initiated by the groundbreaking work of Michael Szwarc reported in the year 1956 in two seminal works1, 2 (Figure 1), a vast variety of polymer ...
  66. [66]
    Ziegler-Natta Polypropylene - an overview | ScienceDirect Topics
    Ziegler-Natta PP refers to polypropylene produced using titanium Ziegler–Natta catalysts, which are typically manufactured as magnesium chloride supported ...
  67. [67]
    From DNA to proteins via the ribosome: Structural insights into the ...
    Apr 13, 2010 · The four major steps during protein synthesis by the ribosome are initiation, elongation, termination and recycling (Figure 1). Briefly, in the ...Introduction · Mrna-Trna Translocation · Peptide Bond Formation
  68. [68]
    Translation factor accelerating peptide bond formation on the ... - NIH
    Ribosomes synthesize proteins in all living cells by catalyzing peptide bond formation. The addition of each new amino acid into the growing peptide occurs in ...Introduction · Ef-P Modifications · Ef-P/eif5a In Bioengineering...
  69. [69]
    homologous enzymes catalyze glycogen synthesis and degradation
    Glycogen and starch synthases are classified in two large protein families: mammalian and yeast GSs (GT3 family) are ∼80 kDa enzymes that use UDP‐glucose as ...
  70. [70]
    From the seminal discovery of proteoglycogen and glycogenin to ...
    Glycogenin-1, GYS, and branching enzyme are the three main enzymes involved in de novo glycogen biogenesis whereas glycogen phosphorylase, debranching enzyme, ...<|separator|>
  71. [71]
    DNA Replication - The Cell - NCBI Bookshelf - NIH
    First, all polymerases synthesize DNA only in the 5′ to 3′ direction, adding a dNTP to the 3′ hydroxyl group of a growing chain. Second, DNA polymerases can add ...
  72. [72]
    Polyhydroxyalkanoates (PHAs) synthesis and degradation by ...
    Sep 1, 2023 · Polyhydroxyalkanoates (PHAs) are promising biodegradable plastics that can be produced by many microbes using various substrates from waste feedstock.
  73. [73]
    Polyhydroxyalkanoates (PHAs): Mechanistic Insights and ... - MDPI
    The pathway in Cupriavidus necator (Ralstonia eutropha) is one of the most thoroughly studied routes for PHA production. In this process (Pathway I), 3- ...Polyhydroxyalkanoates (phas)... · 3. Overview Of Pha... · 5. Biodegradation Of Phas
  74. [74]
    Review on emerging trends and challenges in the modification of ...
    Grafting, cross-linking, curing, blending, and derivatization techniques are used to improve the natural properties of XG and can be used more effectively ...
  75. [75]
    Current advances in the chemical functionalization and potential ...
    Feb 7, 2023 · This review highlights recent progress in the chemical modification of GG and its derivatives based on nucleophilic reactions, partial oxidation, graft ...
  76. [76]
    [PDF] rayon manufacture by the viscose - Tennessee Academy of Science
    In 1891 Cross and Bevan discovered the underlying reaction which led to the development of the viscose process. The product of this reaction is alkali-cellulose ...
  77. [77]
    Cellulose Xanthate - an overview | ScienceDirect Topics
    The third process, which in a variety of forms is now dominant, began commercial production in 1905 as viscose rayon. However, there are environmental problems ...Missing: history | Show results with:history
  78. [78]
    Sulfur Vulcanization - an overview | ScienceDirect Topics
    In 1839, the process of sulfur vulcanization of rubbers was invented by Charles Goodyear and as a result of his discovery the rubber industry was ...
  79. [79]
    “Re-Think” Sulfur Curing - PMC - NIH
    Nov 2, 2024 · Since Charles Goodyear discovered the method of sulfur curing Natural Rubber in 1839, many studies have been carried out to understand its ...
  80. [80]
    Synthesis of hybrid materials using graft copolymerization on non ...
    Graft copolymerization is a common strategy to modify natural polymers and their derivatives with synthetic polymers [3], [19], [20], [21]. A graft ...
  81. [81]
    Polymer Grafting and its chemical reactions - Frontiers
    Graft copolymerization is an efficient way of adding beneficial applications to the major polymer backbone, which is utilized in a variety of uses.Techniques of polymer grafting · Application of graft... · Recent advancements in...
  82. [82]
    Semi-Natural Superabsorbents Based on Starch-g-poly(acrylic acid)
    Aug 10, 2020 · Another set of superabsorbents was obtained by graft polymerization of starch using KPS as an initiating agent (Figure 2). The starch solution ...
  83. [83]
    [PDF] Starch graft copolymers as superabsorbents obtained via reactive ...
    It confirmed starch modification via grafting. Starch graft copolymers with hydrophilic acrylic monomers can be applied as potential water superabsorbents. ...
  84. [84]
    Chitosan as Antimicrobial Agent: Applications and Mode of Action
    Chitosan, a hydrophilic biopolymer industrially obtained by N-deacetylation of chitin, can be applied as an antimicrobial agent. The current review of 129 ...
  85. [85]
    New protocol for the isolation of nitrocellulose from gunpowders
    Introduction. Nitrocellulose, discovered in 1846 by Christian Friedrich Schönbein [1], is a nitrated cellulose ester polymer where the different monomer units ...<|control11|><|separator|>
  86. [86]
    How to Measure the Mechanical Properties of Polymers - AZoM
    Aug 25, 2021 · At its most basic, a tensile test can be used to measure breaking force (N) and elongation at break (mm). ... These can give information on yield ...
  87. [87]
    Tensile Property Testing of Plastics - MatWeb
    Typical Tensile Strength, Elongation, and Tensile Modulus of Polymers ; Nylon 6, 70, 90 ; Polyamide-Imide, 110, 6 ; Polycarbonate, 70, 100 ; Polyethylene, HDPE, 15 ...
  88. [88]
    Polymers - Physical Properties - The Engineering ToolBox
    Densities, tensile strength, elongation, Youngs modulus and Brinell hardness. Physical properties of some common plastic materials are indicated below.
  89. [89]
    Mechanical properties of amorphous and semi-crystalline semi ... - NIH
    Apr 25, 2020 · At 23 °C and 1 Hz, the storage modulus E' is high and is 2.92 GPa for Polyamide A, 3.01 GPa for Polyamide B and 2.96 GPa for Polyamide C, which ...
  90. [90]
    Characteristics of polyurethane (elongation, strength, shock ...
    Compared to other elastomers, it has a elongation at break of about 300 to 700%, and it is superior in shock absorption and durability.
  91. [91]
    Viscoelasticity - an overview | ScienceDirect Topics
    Creep recovery and stress relaxation are the two general tests to define the viscoelastic behavior of polymers. In a creep test, a constant load is applied on ...
  92. [92]
    Constitutive Equations for Analyzing Stress Relaxation and Creep of ...
    May 23, 2022 · The mechanical behaviors of all polymers are viscoelastic, meaning that they exhibit time-dependent mechanical behaviors of both viscous fluids ...
  93. [93]
    Polystyrene PS - detailed analysis - Linseis
    The Glass Transition Temperature (Tg) of Polystyrene typically occurs around 100 °C. This temperature marks the transition of the amorphous polymer from a hard, ...The origin of Polystyrene · The glass transition... · The production of Polystyrene...Missing: source | Show results with:source
  94. [94]
    Glass Transition Temperature-Composition Relationship of ... - Nature
    The Fox equation is based on the assumption that cer- tain properties of a copolymer, e.g., specific volume, mo- lar cohesive energy, or chain stiffness are ...
  95. [95]
    Understanding the Melting Point of High-Density Polyethylene (HDPE)
    Jul 19, 2024 · A: The melting point for High-Density Polyethylene is usually between 120 and 130 degrees Celsius (248 to 266 degrees Fahrenheit). Knowing this ...What is the Typical Melting... · What Are the Environmental...
  96. [96]
    Sensitivity of Polymer Crystallization to Shear at Low and High ...
    Mar 30, 2018 · Flow-induced crystallization (FIC) is a dominant mechanism of polymer self-assembly, but the process is poorly understood at high supercooling ...
  97. [97]
    [PDF] Mean Field Flory Huggins Lattice Theory
    Inspection of solubility parameters can be used to estimate possible compatibility (miscibility) of solvent-polymer or polymer-polymer pairs. This approach ...
  98. [98]
    Thermomechanical Properties of Nontoxic Plasticizers for Polyvinyl ...
    May 11, 2023 · The internally plasticized PVCs show the lower Tg compared to PVC. The Tg of the neat PVC is 84°C, meanwhile, the Tg of the PVC-PDMS, PVC-TBC1, ...
  99. [99]
  100. [100]
    Electrical conductivity of polyaniline (PANI) assisted by anionic ...
    The highest electrical conductivity value obtained is 3.97 S/cm. 1. Introduction. The so–called intrinsically conducting polymers (ICPs) has been type of ...
  101. [101]
    A Comprehensive Review of Piezoelectric PVDF Polymer ... - MDPI
    PVDF is notable for its exceptional piezoelectricity, which refers to its capability to produce an electric charge when a mechanical stress or force is applied, ...
  102. [102]
    What is PMMA? - Gaggione
    This uncrystallized polymer shows remarkable transparency (92% light transmission) in the visible range from 380 to 780 nm. The angle of total reflection on an ...
  103. [103]
    Prediction of birefringence for polymer optical products based on a ...
    Oct 26, 2021 · Actually, the birefringence is essentially caused by the anisotropy of molecular chain polarization related to molecular chain orientation. The ...
  104. [104]
    Refractive Index of Polymers by Index – scipoly.com
    Polymer, Refractive Index. Poly(hexafluoropropylene oxide), 1.3010. Poly(tetrafluoroethylene-co-hexafluoropropylene), 1.3380.
  105. [105]
    and Hansen Solubility Parameters for Various Polymers
    This table highlights polar/dispersion and acid—base breakouts of surface free energy, as well as Hansen Solubility Parameters and radius of interaction, for a ...
  106. [106]
    Polyesters - Chemistry LibreTexts
    Jan 22, 2023 · The polyester forms and half of the ethane-1,2-diol is regenerated. This is removed and recycled. manufpet2.gif. Hydrolysis of polyesters.
  107. [107]
    Polymer OLEDs (PLED): introduction and market status
    Basically you can make OLEDs from two kinds of materials: small-molecule (SM-OLED) or large-molecules, or polymers. Virtually all OLED displays on the market ...
  108. [108]
    [PDF] Plastics – the fast Facts 2024
    ”Plastics – the fast Facts” 2024 shows 2023 preliminary global and European plastics production data. It also provides 2023 European plastics industry's key ...
  109. [109]
    The Global Market of PET Production: from Origins to Recycling
    Therefore, some sources state that in 2024, nearly 28 Mt of PET was produced and China was the main producer with a global share of 31%. Adequate management of ...
  110. [110]
    PVC Pipes Market Size, And Industry Trends Report 2033
    The global PVC pipes market size reached 25.90 Million Tons in 2024 and is estimated to grow 36.3 Million Tons by 2033, at a CAGR of 3.8% from 2025 to 2033.
  111. [111]
    Current progress in synthesis of polyurethane materials based on ...
    Jun 23, 2025 · This article provides an overview of eco-friendly polyurethane materials and additives used in their production.
  112. [112]
    Rigid Polyurethane Foams Market Size, Share and Trends, 2033
    Oct 22, 2025 · The global rigid polyurethane foams market size was valued at USD 22.76 Bn in 2024, and it is expected to reach USD 38.23 Bn by 2033, ...
  113. [113]
    Rubber Industry Platform Aims to Halt Deforestation, Protect Human ...
    The global tire industry buys 70 percent of all natural rubber, meaning it has a clear role in ensuring a sustainable natural-rubber value chain.
  114. [114]
    Global synthetic rubber market forecast at $62 billion by 2037
    Sep 15, 2025 · The global synthetic rubber market was valued at USD 24.29 billion in 2024 and is expected to reach approximately USD 25.75 billion in 2025 ...
  115. [115]
    Automotive Polymer Composites Market Size Report, 2030
    The automotive polymer composites market was valued at USD 10.20 billion in 2023 and is projected to reach USD 14.55 billion by 2030, growing at a CAGR of 5.0%.
  116. [116]
    Materials Market Report 2024 - Textile Exchange
    Sep 26, 2024 · Polyester remained the most produced fiber globally, accounting for 57% of total fiber production. Recycled synthetics face challenges: Although ...
  117. [117]
    Adhesives and Sealants Market Size & Share Report, 2032
    Oct 13, 2025 · The global adhesives and sealants market size is projected to grow from $82.88 billion in 2024 to $118.83 billion by 2032, at a CAGR of 4.6%.
  118. [118]
    By 2060, global production and use of plastics forecasted to triple
    Jul 25, 2024 · By 2060, global plastic use is forecasted to triple to 1,231 Mt, with waste increasing to 1,014 Mt. 50% of waste will be in landfills.
  119. [119]
  120. [120]
    Poly Ethylene Glycol (PEG)‐Based Hydrogels for Drug Delivery in ...
    Apr 13, 2023 · This review article addresses the merits of PEG hydrogels from various perspectives, including controlled drug release, biocompatibility, and ...
  121. [121]
    Recent advances in modified poly (lactic acid) as tissue engineering ...
    Mar 20, 2023 · The advances in modified PLA as tissue engineering materials are discussed in light of its drawbacks, such as biological inertness, low cell adhesion, and low ...
  122. [122]
    Rapid in situ forming PEG hydrogels for mucosal drug delivery
    Apr 3, 2025 · These PEG hydrogels rapidly form in 30 seconds or less, adhere to mucosal tissues, and release protein-based cargoes over several hours.
  123. [123]
    Drug delivery systems based on polyethylene glycol hydrogels for ...
    Jan 29, 2023 · This review provides a theoretical basis and fabrication strategy for the application of PEG-based composite drug delivery systems in local bone defects.
  124. [124]
    Performance comparison of PLA- and PLGA-coated porous ...
    This review paper provides a comprehensive comparison of PLA- and PLGA-coated bioceramic scaffolds which are mainly employed in bone tissue engineering.
  125. [125]
    [PDF] Review of the Potential use of Poly (lactic-co-glycolic acid) as ...
    Mar 23, 2024 · The paper reviews the current development of the biocompatible material PLGA as bone tissue scaffolds. It focuses on the applications, ...
  126. [126]
    Silicone‐based biomaterials for biomedical applications
    May 26, 2021 · Silicone elastomers are widely used in medical device applications. Below are some of the common forms of silicone-based materials. TABLE 1 ...
  127. [127]
    The Use of Silicone in Medicine - AZoM
    Nov 22, 2024 · It is commonly used in breast implants, maxillofacial prosthetics, joint prosthetics, and implantable drug delivery devices. Its flexible nature ...
  128. [128]
    Topical review: Twenty-five years of silicone hydrogel soft contact ...
    Jun 20, 2025 · SiHy materials have largely eliminated hypoxia as a complication seen in contact lens clinical practice, and when used for daily wear, in ...
  129. [129]
    Engineering of mRNA vaccine platform with reduced lipids and ...
    Oct 7, 2025 · The prepared Mn-mRNA nanoparticle is subsequently coated with lipids to form the resulting nanosystem, L@Mn-mRNA, which achieved nearly twice ...
  130. [130]
    Advances in Lipid Nanoparticles for mRNA-Based Cancer ... - Frontiers
    Lipid-polymer hybrid nanoparticles consist of a biodegradable mRNA-loaded polymer core coated with a lipid layer (Persano et al., 2017; Islam et al., 2018 ...<|separator|>
  131. [131]
    Emerging lipid–polymer hybrid nanoparticles for genome editing
    New generations of lipid-polymer hybrid nanoparticles are rapidly emerging as potentially valuable alternatives to upgrade mainstream gene delivery toolboxes.
  132. [132]
    CRISPR/Cas9 Delivery Systems to Enhance Gene Editing Efficiency
    Modified natural polymer carriers, such as chitosan, have also demonstrated significant potential in CRISPR/Cas9 delivery due to their ideal biodegradability ...
  133. [133]
    D638 Standard Test Method for Tensile Properties of Plastics - ASTM
    Jul 21, 2022 · 4.1 This test method is designed to produce tensile property data for the control and specification of plastic materials.
  134. [134]
    ASTM D638: tensile properties plastics - ZwickRoell
    The ASTM D638 standard describes tensile testing on plastics. It is applied to measure tensile properties including the tensile modulus, yield stress, yield ...
  135. [135]
    [PDF] Overview of Glass Transition Analysis by Differential Scanning ...
    This note will discuss the available approaches for the analysis of a glass transition by Differential Scanning Calorimetry (DSC).
  136. [136]
    [PDF] Characterization of Polymers Using TGA
    The TGA results show that the nylon 6,6 polymer undergoes thermal degradation beginning at 482 C and with a total mass loss of 99.0%. There is a small amount ...
  137. [137]
    Understanding Melt Index and ASTM D1238
    Jan 1, 2013 · The MFI is the mass flow rate in a pressure driven flow through a standardized abrupt cylindrical contraction into a short tube performed under ...
  138. [138]
    Electrochemical Impedance Spectroscopy A Tutorial
    This tutorial provides the theoretical background, the principles, and applications of Electrochemical Impedance Spectroscopy (EIS) in various research and ...The Impedance of an Electrical... · Simulation of Impedance Data...
  139. [139]
    Impedance Spectroscopy, AC Conductivity, and Conduction ...
    Jul 14, 2022 · The impedance spectroscopy analysis of these studies showed that the highest ionic conductivity of polymer electrolyte was obtained at 3 wt. % ...
  140. [140]
    The effect of end groups on the thermal degradation of poly(methyl ...
    PMMA begin to degrade slowly at 220 °C, and then 40–47% degrade in the temperature range 220–270 °C, but subsequent heating at 305°C led to 100% degradation.
  141. [141]
    Polymer Degradation - an overview | ScienceDirect Topics
    While the degradation process represents failure of the polymer to perform ... In nature, polymer degradation is induced by thermal activation, hydrolysis ...
  142. [142]
    [PDF] Review of ageing methods & lifetime prediction for polymeric materials
    This report reviews failure mechanisms commonly experienced in polymeric materials and the techniques that can be used to predict their useful service life ...
  143. [143]
    Photodegradation and photostabilization of polymers, especially ...
    UV radiation causes photooxidative degradation which results in breaking of the polymer chains, produces free radical and reduces the molecular weight.
  144. [144]
    Characterization of photo-oxidative degradation process of ...
    Dec 2, 2022 · The photo-oxidative degradation of HDPE, LDPE, LLDPE, and itPP without and with oxo-biodegradable additives induced by UV exposure test was ...
  145. [145]
    Hydrolytic Degradation and Erosion of Polyester Biomaterials - PMC
    Aliphatic polyesters biodegrade via hydrolytically-labile ester bonds. A notable aspect of polyester degradation is the acidic by-products of the hydrolytic ...
  146. [146]
    Hydrolytic Degradation and Erosion of Polyester Biomaterials
    Jul 30, 2018 · Aliphatic polyesters biodegrade via hydrolytically labile ester ... The polymers degraded by hydrolysis and the mechanism was insensitive to pH.Author Information · References
  147. [147]
    Revising the mechanism of polymer autooxidation - RSC Publishing
    Nov 11, 2010 · Recognition of the real mechanism of autooxidation in polymers is a key to developing strategies for the prevention of their degradation.
  148. [148]
    Environmental stress cracking of high-density polyethylene under ...
    Environmental Stress Cracking of HDPE occurred in an aqueous surfactant solution. Environmental Stress Cracking resistance was influenced by the stress state.
  149. [149]
    Plastic Pollution Around the World - Geneva Environment Network
    Aug 8, 2025 · An estimated 14 million metric tons of plastic enters the ocean annually, with estimates of over 170 trillion plastic particles afloat in the ...<|separator|>
  150. [150]
    The degradation potential of PET bottles in the marine environment
    Mar 22, 2016 · Thus, plastics can last in the marine environment for decades or even hundreds of years when in surface; likely far longer when in deep sea.
  151. [151]
  152. [152]
    State of the art on biodegradability of bio-based plastics containing ...
    Several conditions were tested, including aerobic conditions that lead to 90% transformation of PLA into CO2 in compost at 58°C ± 2°C after 70 days, according ...<|separator|>
  153. [153]
    Composting Performance of l-Poly(lactic acid), d ... - ACS Publications
    May 1, 2025 · Poly(lactic acid) (PLA) is a biodegradable polymer derived from renewable resources, which decomposes under industrial composting conditions.
  154. [154]
    The Truth about the Biodegradability of PLA Filament
    Aug 21, 2023 · PLA is only biodegradable under industrial composting conditions and anaerobic digestion – there is no evidence of PLA being biodegradable in soil, home ...
  155. [155]
    Plastics Recycling With Enzymes Takes a Leap Forward - NREL
    Jun 30, 2025 · Advanced Manufacturing Research ... Key Process Improvements Save Energy and Cut Costs for Recycling Polyester With Enzymes. June 30, 2025.
  156. [156]
    Recent advances in enzyme engineering for improved ... - Nature
    Aug 20, 2025 · In the last ~20 years, a multitude of natural enzymes have been discovered that can catalyze the breakdown of the common plastic ...
  157. [157]
    Mandatory recycled content - Plastics Europe
    European plastics producers have called for a mandatory EU recycled content target for plastics packaging of 30% by 2030.
  158. [158]
    Bio-based Polyethylene: A Sustainable Solution for Plastic Waste
    Jul 18, 2024 · Bio-PE is derived from biological sources like sugarcane instead of petroleum. The manufacturing process begins with extracting ethanol from ...
  159. [159]
    I'm green™ bio-based Polyethylene - FKuR
    Green PE is a bio-based polyethylene produced from the renewable raw material sugar cane. As a drop-in, Bio-PE is a regenerative alternative to fossil ...
  160. [160]
    Brief Guide to Polymer Nomenclature
    Monomers can be named using IUPAC recommendations, or well-established traditional names. ... repeating unit (CRU). It can be determined as follows: (i) a ...
  161. [161]
    [PDF] A brief guide to polymer nomenclature (IUPAC Technical Report)
    Mar 22, 2021 · Source-based nomenclature can be used when the monomer can be identified. Alternatively, more explicit structure-based nomenclature can be used ...
  162. [162]
    [PDF] report on nomenclature - dealing with steric regularity - iupac
    Steric order in the main chain can also be designated as tacticity. A tactic polymer is one in which there is an ordered structure with respect to the con-.
  163. [163]
    [PDF] For Peer Review Only - IUPAC
    Dec 23, 2016 · Accordingly, the present document provides recommendations for naming dendrimers and hyperbranched polymers on the basis of structure-based ...
  164. [164]
  165. [165]
    Bakelite® First Synthetic Plastic - American Chemical Society
    Chemists did not fully understand or identify polymers until around 1900. But as early as 1861, the British chemist Thomas Graham had noted that when he ...