Fact-checked by Grok 2 weeks ago

Copolymer

A copolymer is a polymer derived from the polymerization of two or more distinct species, forming a macromolecular chain that incorporates structural units from each , in contrast to a homopolymer which consists of only one type of repeating unit. This process, known as copolymerization, allows for the creation of materials with tailored physical, chemical, and mechanical properties that combine or enhance those of the individual homopolymers. Copolymers are classified based on the arrangement of monomer units along the polymer chain, which significantly influences their behavior and applications. The main types include random copolymers, where monomer units are distributed irregularly; alternating copolymers, featuring a strict ABAB sequence; block copolymers, consisting of long sequential blocks of one monomer type followed by another (e.g., AAAA-BBBB); and graft copolymers, where branches of one homopolymer are attached to the backbone of another. These architectures can lead to unique phenomena, such as microphase separation in block copolymers, enabling into nanostructures. Notable examples of copolymers demonstrate their versatility in industrial and scientific contexts. Styrene-butadiene rubber (SBR), a random copolymer of styrene and 1,3-butadiene, is widely used in due to its elasticity and durability. (ABS), a terpolymer (three monomers), provides high impact resistance and is employed in pipes, automotive parts, and consumer goods. Block copolymers like polystyrene-block-polybutadiene find applications in elastomers, while graft copolymers such as high-impact polystyrene enhance toughness in and . Overall, copolymers play a critical role in , enabling innovations in biomedical devices, adhesives, coatings, and nanocomposites through precise control of monomer composition and .

Fundamentals

Definition and Nomenclature

A copolymer is a polymer derived from more than one species of monomer. In contrast, a homopolymer is derived from only one species of monomer. Copolymers from two monomer species are sometimes called bipolymers, from three terpolymers, and from four quaterpolymers. The composition of a copolymer refers to the relative proportions of the different monomer units, typically expressed as mole fractions or mole percentages. The sequence distribution describes the manner in which these monomer units are arranged along the polymer chain, which can vary from random to more ordered patterns. Mole fractions are indicated in copolymer nomenclature by placing them in parentheses after the name, such as poly(A-co-B) (0.70:0.30 mol/mol), denoting 70 mol% A and 30 mol% B. The specific sequence distribution is influenced by reactivity ratios, which affect the likelihood of one monomer adding to the growing chain relative to another. IUPAC nomenclature for copolymers employs source-based naming, where the prefix "poly" is followed by the monomer names connected by italicized qualifiers that indicate the arrangement. For unspecified sequence, "-co-" is used, as in poly(styrene-co-acrylonitrile); for random, "-ran-"; for block, "-block-"; and for graft, "-graft-". Monomer names are alphabetized within the name unless a specific sequence must be preserved. The first major synthetic copolymer, rubber (Buna S), was developed in in 1929 through copolymerization as a substitute for during shortages. A common example is styrene- copolymer (SAN), a transparent typically containing 70-75% styrene and 25-30% acrylonitrile, valued for its rigidity and chemical resistance in applications like housings and appliances.

Copolymerization Mechanisms

Copolymerization primarily occurs through two broad classes of mechanisms: chain-growth and . In chain-growth copolymerization, the process involves the sequential addition of monomers to an active chain end, typically initiated by species that generate reactive centers on the monomers. This mechanism is widely used for synthesizing copolymers from or olefinic monomers and proceeds via radical, anionic, cationic, or coordination pathways. Free radical chain-growth copolymerization is one of the most common methods, initiated by thermal or photochemical decomposition of initiators such as peroxides to form radicals that add to the double bond of a monomer, creating a propagating radical chain end. Propagation occurs through rapid addition of subsequent monomers, with the rate governed by the equation for monomer 1 consumption: -\frac{d[M_1]}{dt} = k_{p1} [M_1][R^\bullet] where k_{p1} is the propagation rate constant, [M_1] is the concentration of monomer 1, and [R^\bullet] is the concentration of the propagating radical; this extends analogously to the second monomer and cross-additions in copolymerization. Anionic copolymerization, suitable for monomers with electron-withdrawing groups like styrene or acrylonitrile, begins with nucleophilic initiators such as organolithium compounds that deprotonate or add to the monomer, forming a carbanion that propagates by attacking additional monomers. Cationic copolymerization, employed for monomers stabilizing carbocations such as isobutylene, uses initiators like proton acids (e.g., BF₃ with water co-initiator) to generate a carbocation chain end that adds monomers sequentially. Coordination copolymerization, often for olefins like ethylene and propylene, relies on transition metal catalysts such as Ziegler-Natta systems (e.g., TiCl₄ with AlR₃), where monomers coordinate to the metal center before insertion into the growing chain, enabling stereoregular copolymers under controlled temperature and pressure conditions. Step-growth copolymerization, in contrast, involves the of bifunctional monomers to form covalent bonds between functional groups, typically through with elimination of small molecules like . This mechanism is prevalent for producing copolymers such as polyesters from diols and diacids (e.g., from and ) or polyamides from diamines and diacids (e.g., nylon 6,6 from and ). The process requires stoichiometric balance of monomers and often elevated temperatures (e.g., 200–300°C) under reduced to drive toward high molecular weight, with catalysts like acids or bases accelerating the steps. In both mechanisms, the incorporation of monomers into the copolymer is influenced by factors such as differences in monomer reactivity, which determine the preference for homopropagation versus cross-propagation, and , which affects monomer availability in the medium— for instance, poor solubility of one monomer can lead to heterogeneous incorporation and phase-separated domains in the resulting copolymer.

Reactivity Ratios

Reactivity ratios are fundamental kinetic parameters in copolymerization that govern the relative rates at which different monomers are incorporated into the growing polymer chain, thereby influencing the sequence distribution. The reactivity ratio for monomer 1, denoted r_1, is defined as the ratio of the propagation rate constant for addition of monomer 1 to a chain-end radical derived from monomer 1 (k_{11}) to the rate constant for addition of monomer 2 to the same radical (k_{12}), expressed as r_1 = \frac{k_{11}}{k_{12}}. Similarly, the reactivity ratio for monomer 2 is r_2 = \frac{k_{22}}{k_{21}}, where k_{21} and k_{22} are the corresponding rate constants for the chain-end radical derived from monomer 2. The Mayo-Lewis equation relates the instantaneous of 1 incorporated into the copolymer (F_1) to the monomer feed composition (f_1 and f_2 = 1 - f_1) through the reactivity ratios: F_1 = \frac{r_1 f_1^2 + f_1 f_2}{r_1 f_1^2 + 2 f_1 f_2 + r_2 f_2^2} This equation, derived from steady-state assumptions in free radical copolymerization, predicts composition drift as proceeds if r_1 \neq r_2. The magnitudes of r_1 and r_2 interpret the copolymerization behavior: when both equal 1, the system exhibits ideal random incorporation, with copolymer composition matching the feed. Values where r_1 > 1 and r_2 < 1 (or vice versa) indicate that each radical preferentially adds its own monomer, promoting gradient or block-like sequences. Conversely, both r_1 < 1 and r_2 < 1 signify a preference for cross-addition, favoring alternating structures. Experimental determination of reactivity ratios typically involves copolymerizations at low monomer conversions (<10%) to ensure the instantaneous composition approximates the feed, followed by analysis of copolymer composition via techniques such as nuclear magnetic resonance (NMR) spectroscopy for sequence assignment or gravimetric/elemental methods for overall monomer content. Data from multiple feed compositions are then fitted to the using nonlinear least-squares optimization to yield r_1 and r_2; for processes at higher conversions, numerical integration of the differential copolymer equation accounts for evolving feed composition. Illustrative examples highlight diverse behaviors: in the free radical copolymerization of styrene (monomer 1) and methyl methacrylate (monomer 2) at 60°C, r_1 \approx 0.52 and r_2 \approx 0.46, reflecting nearly ideal random copolymerization with mild alternation tendency. By contrast, the butadiene (monomer 1)-styrene (monomer 2) system in certain anionic conditions yields r_1 \approx 0.1 and r_2 \approx 5, exhibiting strong block-forming propensity due to the butadienyl radical's high self-preference.

Linear Copolymers

Alternating Copolymers

Alternating copolymers feature a highly ordered linear structure in which two distinct monomer units, denoted as A and B, alternate strictly along the chain in a repeating ...ABABAB... sequence. This regularity stems from copolymerization kinetics where the reactivity ratios r_1 (for monomer A adding to its own radical) and r_2 (for monomer B adding to its own radical) are both less than 1, thereby disfavoring homopropagation and strongly preferring cross-propagation steps. Such conditions ensure a near-ideal 1:1 incorporation ratio, independent of the initial monomer feed composition. The alternation is particularly pronounced in systems involving electron-donor and electron-acceptor monomers, where charge-transfer complexes form between the comonomers, stabilizing the propagating radical and directing selective addition. For instance, styrene serves as the donor while maleic anhydride acts as the acceptor, enabling the synthesis of alternating copolymers through conventional free radical polymerization initiated by agents like or at temperatures of 60–80°C. In this exemplary system, the reactivity ratios are approximately r_1 \approx 0.01 for styrene and r_2 \approx 0.03 for maleic anhydride, resulting in copolymers with exceptional sequence control and molecular weights often exceeding 10^5 g/mol. These copolymers exhibit superior thermal stability, with decomposition temperatures typically above 300°C, attributed to the uniform distribution of functional groups that hinders chain unzipping compared to random copolymers. Additionally, their balanced polarity enhances compatibility in polymer blends, facilitating better phase dispersion and mechanical integrity. Common applications include adhesives and coatings, where the alternating structure provides strong adhesion to diverse substrates and resistance to environmental degradation; however, their utility remains confined to monomer pairs with inherent alternation proclivity, such as donor-acceptor combinations.

Periodic Copolymers

Periodic copolymers are linear macromolecules featuring a predefined, repeating sequence of multiple distinct monomeric units along the chain, typically involving three or more species arranged in a regular pattern. This structure is exemplified by repeating blocks such as (ABC)_m, where A, B, and C denote different monomers, distinguishing them from simpler binary repetitions. The feasibility of achieving such regularity is influenced by the reactivity ratios of the monomers, which govern the propensity for cross-propagation over homopolymerization. Unlike alternating copolymers limited to two-unit (AB)_n sequences, periodic copolymers enable more complex motifs with longer repeating segments, such as ordered arrangements of , , and units. These are synthesized via controlled living polymerization techniques, particularly with sequential monomer addition, allowing precise programming of the sequence. Macroinitiators further facilitate this process by initiating polymerization of subsequent units in a controlled manner, yielding high molecular weight polymers with narrow polydispersity. The periodic arrangement imparts tunable crystallinity and mechanical strength, often surpassing those of random or block analogs. For instance, poly(ethylene-per-ethylene-per-methyl methacrylate) exhibits approximately 30% crystallinity and a melting temperature of 90°C, despite amorphous statistical counterparts, with a Young's modulus of 10³ kg/cm² and elongation at break of 300%. This ordered structure enhances overall durability and elasticity. Representative examples include periodic styrene-isoprene copolymers (St-per-Is), prepared through living anionic polymerization with programmed monomer dosing, which display glass transition temperatures exceeding those of random styrene-isoprene variants, supporting applications in elastomers. Such periodic copolymers are incorporated into synthetic rubber formulations for tire production, where the sequence control optimizes mechanical performance like wear resistance and grip.

Statistical Copolymers

Statistical copolymers are linear polymers in which the monomer units are distributed randomly along the chain, with the sequence of units following probabilistic laws rather than a regular pattern. This random distribution can be modeled using for independent monomer additions or to account for dependencies on the previous unit, enabling the calculation of probabilities for specific sequences such as dyads or triads. In ideal cases, the incorporation of each monomer type occurs with equal probability regardless of the chain end, resulting in a uniform statistical randomness across the polymer. These copolymers form primarily through free radical or other chain-growth polymerization mechanisms when the reactivity ratios r_1 and r_2 of the two monomers are approximately equal to 1, indicating similar reactivities toward both types of propagating radicals. In near-ideal scenarios, this leads to a random copolymer composition that matches the monomer feed ratio at low conversions, but as polymerization proceeds to higher conversions, a composition drift occurs due to differential monomer consumption. The average composition of the copolymer at finite conversions is described by the integrated form of the , which relates the overall copolymer mole fraction to the initial and final monomer feed compositions, accounting for this drift through numerical integration or approximation. For non-ideal randomness, reactivity ratios slightly deviating from unity introduce subtle biases in sequence distribution without forming ordered structures. On a large scale, statistical copolymers appear homogeneous, with properties averaging those of the constituent homopolymers, but they exhibit microheterogeneity due to local variations in sequence lengths and compositions along individual chains. This microheterogeneity can enhance miscibility in polymer blends by reducing interfacial tensions and promoting compatible mixing at the molecular level. A representative example is poly(styrene-co-acrylonitrile) (SAN), a statistical copolymer typically containing 25-30% acrylonitrile, which improves impact resistance, chemical stability, and dimensional stability compared to polystyrene while maintaining transparency and rigidity. SAN is widely used in applications requiring enhanced mechanical toughness, such as housings and automotive components.

Gradient Copolymers

Gradient copolymers are a class of linear copolymers characterized by a gradual variation in monomer composition along the polymer chain, typically transitioning from one monomer-rich end to another. This structure arises naturally in free radical copolymerizations of monomers with differing reactivities (reactivity ratios r₁ ≠ r₂), particularly at low conversions where the instantaneous monomer feed composition shifts as the more reactive monomer is preferentially incorporated early in the reaction. For instance, in systems like methyl methacrylate (MMA) and n-butyl acrylate (BA), compositional drift results in chains that are MMA-rich at the initiating end and BA-rich toward the terminating end. To achieve precise control over the gradient profile, advanced synthesis methods employ controlled radical polymerization techniques such as atom transfer radical polymerization (ATRP), nitroxide-mediated polymerization (NMP), and reversible addition-fragmentation chain transfer (RAFT) polymerization. These living/controlled approaches enable uniform chain growth and tunable gradient steepness. Additionally, semi-batch feeding strategies, where monomers are added incrementally during polymerization, allow for deliberate manipulation of the composition drift, producing well-defined gradients even with conventional initiators. A notable example is the use of concurrent tandem living radical polymerization with in situ monomer transformation to generate gradient copolymers in a one-pot process. The gradual compositional change imparts unique properties to gradient copolymers, distinguishing them from random or block architectures. They exhibit broad glass transition temperatures due to the sequential domains of varying composition and, crucially, superior interfacial activity that enhances compatibilization in immiscible polymer blends. By distributing comonomer units in a tapered manner, gradient copolymers localize at interfaces more effectively than diblock copolymers, significantly reducing interfacial tension—for example, symmetric gradient copolymers with linear composition profiles can lower tension in polystyrene-poly(methyl methacrylate) blends by factors exceeding those of blocks with sharp junctions. This makes them valuable for applications requiring improved blend morphology and mechanical properties without phase segregation. Representative examples include gradient poly(MMA-co-BA) copolymers, synthesized via ATRP or semi-batch emulsion polymerization, which leverage their tunable softness gradient for pressure-sensitive adhesives with enhanced tack and peel strength. Similarly, epoxy-functional gradient poly(glycidyl methacrylate-co-n-butyl acrylate) copolymers, prepared by ATRP, demonstrate improved adhesion in composite materials due to their interfacial bridging capabilities.

Block Copolymers

Block copolymers are linear macromolecules composed of two or more chemically distinct homopolymer segments, or blocks, covalently linked in a sequential manner, typically forming diblock (A-B) or triblock (A-B-A) structures. These blocks consist of long chains of the same monomer type, enabling distinct physical behaviors within a single polymer chain, unlike random or alternating copolymers where monomers are interspersed irregularly. The length and composition of each block can be precisely tailored to influence overall properties, with common notations such as polystyrene-block-polybutadiene (PS-b-PBD) illustrating a diblock where polystyrene (A) is followed by polybutadiene (B). Synthesis of block copolymers primarily relies on sequential polymerization techniques that allow controlled addition of monomers, often using living polymerization methods to achieve well-defined structures with narrow molecular weight distributions. Living anionic polymerization, a seminal approach developed in the mid-20th century and refined through subsequent advancements, exemplifies this process: it begins with the anionic polymerization of one monomer (e.g., styrene) using an initiator like sec-butyllithium, followed by the addition of a second monomer (e.g., butadiene) to form the second block without termination. This method provides excellent molecular weight control, with polydispersity indices (PDI) as low as 1.04, due to the absence of chain transfer or termination reactions, allowing block lengths to be adjusted by monomer-to-initiator ratios. In conventional free radical copolymerization, block formation is favored when reactivity ratios (r1 and r2) are both greater than 1, promoting homopolymerization-like sequences of each monomer before crossover, though sequential living methods are preferred for precise control. The defining property of block copolymers arises from the thermodynamic immiscibility of the constituent blocks, which drives microphase separation into ordered domains on the nanoscale, even when the overall material is macroscopically homogeneous; this behavior is elaborated further in discussions of phase morphology. A prominent example is the styrene-butadiene-styrene (SBS) triblock copolymer, synthesized via living anionic polymerization, which serves as a thermoplastic elastomer: the central polybutadiene block imparts rubbery elasticity, while the terminal polystyrene blocks provide thermoplastic processability and mechanical strength through physical cross-linking at room temperature. SBS materials exhibit reversible deformation with elongations up to 900% and recovery, making them widely used in adhesives, footwear, and sealants, with molecular weights typically in the range of 100,000–200,000 g/mol to optimize phase separation and performance.

Stereoblock Copolymers

Stereoblock copolymers are linear homopolymers featuring sequential blocks of the same repeating unit but with differing stereoregularities, typically alternating between isotactic and syndiotactic (or atactic) segments. This configuration arises from controlled variations in the spatial arrangement of substituents along the chain, enabling distinct physical behaviors within a single polymer type. Unlike copolymers defined by constitutional differences in monomer units, stereoblock variants leverage tacticity to form microphase-separated domains that influence overall material performance. Tacticity is characterized using dyad notation in nuclear magnetic resonance (NMR) spectroscopy, where mm denotes meso (isotactic) dyads, rr denotes racemo (syndiotactic) dyads, and mr denotes heterotactic dyads, with fractions reflecting the relative proportions of each configuration. Synthesis of stereoblock copolymers primarily relies on coordination polymerization mechanisms that allow mid-reaction modulation of catalyst stereoselectivity to generate tacticity contrasts. Ziegler-Natta catalysts, variants of which can switch between isospecific and syndiospecific active sites, enable the production of isotactic-syndiotactic blocks by altering reaction conditions or catalyst composition during propagation. Living Ziegler-Natta systems provide enhanced control, facilitating programmable stereomodulation for multiblock isotactic-atactic stereoblock polypropylenes with precise block lengths and narrow molecular weight distributions. Metallocene catalysts with fluxional ligands represent a key advancement, oscillating between chiral and achiral coordination geometries to yield stereoblock architectures without interrupting the reaction. A prototypical example is bis(2-phenylindenyl)zirconium dichloride activated by methylaluminoxane, which polymerizes propylene into isotactic-atactic stereoblock chains; the isotactic pentad content (mmmm) varies from 6.3% to 28.1%, tunable by decreasing temperature or increasing monomer pressure to optimize block contrast. Temperature shifts in living polymerization offer another route, transitioning from syndiospecific conditions at low temperatures (e.g., 0 °C) to isospecific at higher temperatures (e.g., 25 °C), thereby forming isotactic-syndiotactic stereoblock polypropylene with defined junctions observable via NMR. These structural features impart superior mechanical properties to stereoblock copolymers relative to fully atactic homopolymers, including enhanced elasticity from reversible crystallization of tactic blocks within an amorphous matrix. Stereoblock polypropylene exemplifies this, exhibiting thermoplastic elastomeric behavior with elongations up to 1000% and rapid recovery, attributed to the flip-flop reorientation of crystallites in compliant atactic segments, which improves processability and toughness over brittle atactic variants. Chiral initiators in living polymerization extend this approach to other monomers, such as stereoblock poly(lactic acid) produced via ring-opening of racemic lactide, where alternating D- and L-blocks enhance hydrolytic stability and mechanical strength. Applications of stereoblock polypropylene leverage its elastomeric traits for demanding uses, such as flexible components requiring durability and recyclability. Stereoblock poly(lactic acid), with its biocompatible and biodegradable nature, serves as a representative example for biomedical implants, offering tunable modulus and reduced brittleness compared to homopolylactides while maintaining sufficient crystallinity for structural integrity.

Branched and Complex Copolymers

Graft Copolymers

Graft copolymers are a class of branched macromolecules characterized by a linear backbone polymer composed of one monomer type, such as (component A), onto which side chains or grafts of a different monomer type, such as (component B), are covalently attached at various points along the backbone. This structure introduces asymmetry and branching, distinguishing graft copolymers from linear copolymers by enabling unique rheological behaviors, such as reduced viscosity in melts due to the side chains restricting entanglement. The synthesis of graft copolymers primarily employs three strategies: "grafting from," "grafting to," and "grafting through," each offering control over graft density and length depending on the desired architecture. In the "grafting from" method, reactive sites are first introduced onto the preformed backbone—often via chemical modification or irradiation—followed by polymerization of the graft monomer directly from these sites, as exemplified by atom transfer radical polymerization (ATRP) of methyl methacrylate from a functionalized poly(ε-caprolactone) backbone to yield poly(ε-caprolactone)-g-poly(methyl methacrylate) with molar masses around 29,000 g/mol. The "grafting to" approach couples pre-synthesized graft chains to complementary functional groups on the backbone, typically through efficient reactions like copper-catalyzed azide-alkyne cycloaddition (CuAAC), such as attaching azido-terminated polystyrene to alkyne-functionalized poly(ε-caprolactone). Meanwhile, "grafting through" involves copolymerizing backbone monomers with macromonomers that already bear polymerizable end-groups, incorporating the side chains during the main chain formation, as seen in the ring-opening polymerization of ε-caprolactone with methoxy poly(ethylene glycol)-substituted caprolactone to produce amphiphilic grafts for nanoparticle coatings. Key structural parameters in graft copolymers include the degree of grafting (DG), defined as the percentage of backbone repeat units bearing a side chain or the weight ratio of grafted material to the original backbone, and the graft length, measured as the degree of polymerization (DP) of the side chains. These metrics profoundly affect material properties; for instance, in poly(methyl methacrylate)-g-oligo(2-ethyl-2-oxazoline) systems, a DG ranging from 9% to 34% modulates hydrophilicity and cloud point temperature, while side chain DPs of 5 to 24 influence micelle formation and aggregation numbers (around 10 for longer chains), enabling applications like drug solubilization. Higher DG values enhance phase compatibility, but excessive grafting can lead to steric hindrance during synthesis in "grafting to" methods. Graft copolymers impart enhanced mechanical toughness through energy dissipation mechanisms facilitated by the branched architecture, achieving up to 280% strain at break in polylactide blends compared to 7% for unmodified polylactide. They also serve as effective compatibilizers in immiscible polymer blends, reducing interfacial tension and improving adhesion, as demonstrated by polystyrene-g-poly(methyl methacrylate) additives that boost impact strength, tensile strength, and bending strength in polystyrene/poly(methyl methacrylate) mixtures. A representative commercial example is the acrylonitrile-butadiene-styrene (ABS) copolymer, where polystyrene-co-acrylonitrile (SAN) grafts are attached to a polybutadiene rubber backbone via seeded semibatch emulsion polymerization. In this process, a polybutadiene latex seed is swollen with styrene and acrylonitrile monomers in an aqueous emulsion, initiated by a water-soluble radical source, yielding a core-shell morphology with a polybutadiene core (61 vol%) and SAN shell (39 vol%), alongside an internal grafting degree of approximately 7.3% and grafting efficiency of 13.3%. This structure confers high impact resistance and processability to ABS resins used in automotive and consumer goods. Recent progress as of 2025 has focused on graft copolymers in biomedical applications, including enhanced bioadhesives via graft copolymerization for tissue engineering and smart hydrogels for controlled drug release.

Star Copolymers

Star copolymers are a class of branched macromolecules consisting of multiple linear polymer arms radiating from a central core, typically with three or more arms to distinguish them from simple branched structures. The core can be a small multifunctional molecule, a cross-linked oligomer, or a nanoparticle, while the arms may be homopolymer chains or copolymer segments of one or more monomer types, enabling diverse architectures such as symmetric homoarm stars (e.g., all arms identical) or asymmetric miktoarm stars (e.g., A₃B₃ with three A-type and three B-type arms). This radial geometry contrasts with graft copolymers, which attach branches sequentially along a linear backbone. Synthesis of star copolymers employs two primary strategies: the core-first (divergent) approach, where polymerization initiates from a multifunctional core to grow arms outward, and the arm-first (convergent) approach, where preformed linear arms are coupled to a central core. In the core-first method, techniques like (ATRP) or living anionic polymerization use initiators with multiple reactive sites, such as pentaerythritol-based cores for four-arm stars. For miktoarm variants, cross-linking agents like form a (PDVB) core during copolymerization with monomers for different arms, as demonstrated in the synthesis of polyethylene-polystyrene miktoarm stars via anionic polymerization followed by hydrogenation. The arm-first method often involves or coupling reactions to attach arms to cores like cyclotriphosphazene or divinylbenzene-derived moieties, offering high yields and precise control over arm attachment. Star copolymers exhibit distinct physical properties arising from their compact, globular shape, including lower solution viscosity and reduced hydrodynamic volume compared to linear polymers of equivalent molecular weight. This compactness stems from the high segmental density near the core, leading to a branching factor g' = \frac{[\eta]_{\text{star}}}{[\eta]_{\text{linear}}} < 1, where [\eta] is the intrinsic viscosity, often quantified as 0.5–0.7 for typical stars. The number of arms, typically 3–20 for well-defined stars, and their functionality (e.g., end-group reactivity) can be precisely controlled by selecting cores with defined valency and polymerization conditions, influencing chain entanglement and self-assembly behavior. Representative examples include miktoarm star copolymers like A₃B₃ architectures, which enable tailored microphase separation due to differing arm compatibilities. Star poly(ethylene oxide) (PEO) polymers, synthesized via core-first anionic polymerization from cores like 1,1,1-tris(hydroxymethyl)ethane, demonstrate enhanced solubility and biocompatibility, with applications in drug delivery systems where their compact form improves circulation time and payload capacity. As of 2025, star copolymers have seen developments in stimuli-responsive networks for drug delivery and optoelectronic applications, such as star block copolymers enabling tunable charge transport in organic electronics.

Dendrimer-Like Copolymers

Dendrimer-like copolymers are highly branched macromolecules characterized by a tree-like, generational structure originating from a central core, where copolymer segments form the branching units across multiple layers. Unlike traditional dendrimers composed of small monomeric repeats, these copolymers incorporate polymeric chains, such as polystyrene or poly(methyl methacrylate), between branch points, enabling precise control over architecture while maintaining monodispersity with polydispersity indices below 1.05. A representative example involves polyamidoamine (PAMAM) dendrimers as the core with grafted vinyl polymers like polystyrene arms grown via atom transfer radical polymerization (ATRP), creating hybrid structures that combine the rigidity of dendrimer scaffolds with the flexibility of polymer branches. Synthesis of dendrimer-like copolymers typically employs divergent growth strategies through iterative cycles of polymerization and coupling/activation steps, often starting from a multifunctional core. Living/controlled polymerization techniques, such as or , allow for the sequential addition of generations, where each layer involves growing linear copolymer arms (e.g., block copolymers of ) and linking them to peripheral functional groups on the previous generation. This core-first approach has enabled the construction of structures up to the seventh generation, with molecular weights reaching several million g/mol, as demonstrated in the work of and colleagues using iterative divergent methodologies on . Star copolymers can serve as low-generation analogs in this synthesis paradigm, providing initial branched scaffolds for further generational expansion. These copolymers exhibit precise size control and high peripheral functionality, making them ideal for encapsulation applications, where the dense branching creates internal voids capable of hosting guest molecules. The generation number and branching density significantly influence solubility: higher generations increase the density of terminal hydrophilic groups (e.g., in ), enhancing water solubility and reducing aggregation in aqueous media, while excessive branching can induce conformational contraction that limits solvent penetration into the core. For instance, studies on show that solubility improves exponentially with generations beyond the third due to amplified surface amphiphilicity. A notable example from recent advancements is the dendrimer-like star block copolymer composed of poly(ε-caprolactone) (PCL) inner blocks and poly(ethylene glycol) (PEG) outer blocks, synthesized via ring-opening polymerization followed by "click" chemistry for generational linking, achieving up to three generations with enhanced biocompatibility. These structures form stable nanocarriers for drug delivery, leveraging the hydrophobic PCL core for encapsulation and hydrophilic PEG corona for prolonged circulation, as evidenced in applications stabilizing gold nanoparticles and potential theranostic uses in the 2020s. In 2024-2025, dendrimer-like copolymers have advanced in cancer therapy and sensor technologies, including dendrimer-based nanogels for targeted drug delivery and electrochemical sensors for biomarker detection.

Physical Properties

Microphase Separation

Microphase separation in copolymers arises primarily from the incompatibility between distinct polymer segments, quantified by the Flory-Huggins interaction parameter χ, which measures the enthalpic penalty for mixing unlike segments. When χ exceeds a critical value relative to the degree of polymerization N—specifically, when the product χN surpasses approximately 10.5 for symmetric diblock copolymers—the system transitions from a disordered state to ordered nanoscale domains, as predicted by random phase approximation theory. This driving force stems from the unfavorable mixing enthalpy outweighing the entropic cost of segregation, yet the covalent linkage between blocks prevents macroscopic phase separation, confining domains to microscale features. In block copolymers, which primarily enable this self-assembly, the resulting morphologies depend on the volume fraction f of one block relative to the total. For instance, lamellar structures form near f ≈ 0.5, where alternating layers of each block stack periodically; cylindrical domains emerge at f ≈ 0.2–0.4 or 0.6–0.8, with one block forming rods in a matrix of the other; and spherical domains appear at extreme compositions like f ≈ 0.1 or 0.9, yielding micelles of the minority block. These periodic nanostructures typically span 10–100 nm in domain size, reflecting the balance of interfacial energy and chain stretching in the strong segregation regime. The scale and stability of these domains are influenced by molecular weight (via N), block composition (f), and temperature, which modulates χ (often decreasing with rising temperature due to enhanced conformational entropy). Higher molecular weights enlarge domains proportionally to N^{2/3} in mean-field approximations, while shifts in f alter morphology transitions, and temperature changes can induce order-disorder transitions or reorient domains. A representative example is the diblock copolymer of polystyrene (PS) and polybutadiene (PB), where compositions favoring PB cylinders (f_PB ≈ 0.3) yield hexagonally packed cylindrical domains of PB in a PS matrix, demonstrating robust microphase separation for applications in nanostructured materials.

Phase Behavior and Morphology

The phase behavior of copolymers is primarily dictated by the interplay between enthalpic interactions, quantified by the Flory-Huggins parameter χ, and entropic contributions from chain connectivity and architecture, with the dimensionless product χN serving as the key metric for phase stability, where N is the total degree of polymerization. In diblock copolymers, mean-field theory predicts a second-order order-disorder transition (ODT) from a homogeneous disordered state to an ordered microphase-separated morphology. For symmetric diblocks (volume fraction f = 0.5), the critical value is χN_c = 10.495; more generally, for asymmetric compositions, χN_c is higher and typically ranges from 15 to 40 depending on the deviation from f = 0.5, as determined numerically from the random phase approximation within self-consistent field theory. This transition marks the onset of composition fluctuations that grow unstable, leading to periodic nanostructures. In highly incompatible copolymer systems where χN ≫ χN_c, the covalent bonding between dissimilar segments enforces microphase separation on the nanoscale, forming domains such as lamellae or cylinders, in stark contrast to immiscible homopolymer blends that undergo macro-phase separation into large, bulk domains due to the absence of connectivity. This distinction arises because chain tethering in copolymers suppresses long-range diffusion, stabilizing finite domain sizes even at strong segregation. Polymer architecture further modulates this behavior: branched copolymers, including star and graft variants, exhibit suppressed phase separation relative to linear analogs, with elevated χN_c values (often 1.5–2 times higher) stemming from topological constraints that enhance mixing entropy and hinder domain formation. Morphology in copolymer melts evolves dynamically under processing conditions, such as thermal annealing or applied shear, which drive the system toward thermodynamic equilibrium or kinetically trapped states. Annealing facilitates defect annihilation and domain coarsening by enabling chain diffusion, often transforming disordered or polycrystalline structures into long-range ordered phases over timescales proportional to N^2. Shear, conversely, imposes directional forces that align domains parallel to the flow or induce transitions (e.g., from spheres to cylinders), with the extent depending on shear rate and copolymer composition; low rates promote orientation, while high rates can trigger shear-induced disordering above the ODT. Gradient copolymers provide a notable example of tailored phase behavior, where the monotonic composition drift along the chain leads to interfacial widening in immiscible polymer blends. Unlike block copolymers that localize sharply at interfaces, gradients distribute across the boundary, effectively broadening it by 10–50 nm depending on gradient steepness, thereby enhancing compatibilization and suppressing coalescence without forming distinct microphases.

Characterization Methods

Spectroscopic Techniques

Nuclear magnetic resonance (NMR) spectroscopy is a primary tool for determining the composition and sequence distribution in copolymers at the molecular level. Proton (¹H) and carbon-13 (¹³C) NMR spectra provide quantitative information on monomer ratios by integrating peak areas corresponding to specific protons or carbons from each comonomer unit. For instance, in ethylene-propylene copolymers, ¹³C NMR distinguishes methylene and methine carbons to calculate comonomer content and dyad/triad sequences, reflecting copolymerization reactivity ratios. These reactivity ratios, validated through NMR-derived sequence probabilities, enable prediction of microstructural heterogeneity. Infrared (IR) and Raman spectroscopy complement NMR by identifying and quantifying functional groups based on characteristic vibrational modes. The carbonyl stretching band around 1730 cm⁻¹ in acrylate copolymers, for example, allows estimation of acrylate content through peak intensity calibration against reference standards. Raman spectroscopy similarly detects C=O vibrations but is advantageous for aqueous or opaque samples, as seen in monitoring emulsion copolymerizations where it quantifies monomer conversion via band ratios. Both techniques confirm the presence of specific moieties, such as ester groups in poly(acrylic acid)-based copolymers, without requiring sample dissolution. Ultraviolet-visible (UV-Vis) spectroscopy is particularly useful for copolymers incorporating chromophores, such as in conjugated systems where π-π* transitions reveal electronic structure and composition. In donor-acceptor conjugated copolymers like those with carbazole and pentaphenylene units, absorption maxima in the 400-500 nm range track the incorporation of chromophoric segments and their conjugation length. This method aids in assessing optical properties tied to sequence distribution in semiconducting materials. Quantitative analysis via end-group determination, often using ¹H NMR, estimates number-average molecular weight (Mₙ) by comparing end-group signals to repeating unit peaks. For end-functionalized polymers like poly(methyl methacrylate), integration of initiator-derived protons against backbone methoxy signals yields precise Mₙ values, especially for lower molecular weights below 10,000 g/mol. This approach is essential for controlled polymerization techniques where end-group fidelity impacts properties. A representative example is the use of ¹³C NMR to analyze comonomer distribution in styrene-acrylonitrile copolymers, where specific carbon resonances such as methine and quaternary carbons differentiate dyad and triad sequences. Such spectral assignments enable detailed microstructural mapping, crucial for tailoring copolymer performance.

Microscopic and Scattering Methods

Microscopic and scattering methods are essential for visualizing and quantifying the nanoscale morphology of copolymers, particularly their domain structures formed during microphase separation. These techniques provide direct evidence of features such as lamellae, cylinders, and spheres, confirming microphase separation in block copolymers. Transmission electron microscopy (TEM) and atomic force microscopy (AFM) enable high-resolution imaging of domains, while small-angle X-ray scattering (SAXS) and wide-angle X-ray scattering (WAXS) offer statistical insights into periodic structures without destructive sampling. In electron microscopy, TEM is widely used to image copolymer domains at resolutions down to 1 nm, revealing contrasts between immiscible blocks through selective staining. For instance, osmium tetroxide (OsO₄) vapor staining preferentially binds to unsaturated bonds in polyisoprene blocks of polystyrene-block-polyisoprene copolymers, enhancing electron density and highlighting lamellar or cylindrical morphologies in thin sections. Atomic force microscopy (AFM) complements TEM by providing surface topography and mechanical property maps of copolymer films in ambient conditions, with resolutions around 10 nm for domain features in star-shaped or triblock copolymers. AFM's phase imaging mode distinguishes soft and hard domains based on adhesion or modulus differences, as demonstrated in studies of polystyrene-block-polybutadiene-block-polystyrene films where hexagonal cylinder orientations were tracked in real time during annealing. Sample preparation is critical for both techniques due to resolution limits imposed by beam damage and section thickness; TEM typically requires sections thinner than 100 nm to minimize scattering artifacts, while AFM is limited by tip radius to features larger than 5 nm. Ultramicrotomy, using a diamond knife at cryogenic temperatures, produces uniform thin sections from bulk copolymer samples, enabling cross-sectional views of oriented domains without deformation. Selective etching, such as plasma or chemical treatments, further enhances contrast by removing one block, exposing underlying morphologies in etched polystyrene-block-polymethylmethacrylate films for clearer TEM visualization. Scattering methods like SAXS probe domain spacings in the 1-100 nm range by measuring intensity as a function of scattering vector q, where the primary peak position q* corresponds to the characteristic period via the relation q^* = \frac{2\pi}{d}, with d representing the lamellar repeat distance. WAXS extends this to atomic-scale features, such as chain packing in crystalline domains of semicrystalline copolymers. These techniques are non-destructive and suitable for in situ studies, providing ensemble-averaged data on morphology evolution under temperature or shear. Quantitative analysis from these methods includes domain size distributions derived from Fourier transforms of TEM/AFM images or SAXS peak widths, which quantify polydispersity in lamellae spacings of diblock copolymers. Orientation factors, calculated from azimuthal intensity spreads in 2D SAXS patterns, assess alignment degrees, with values near 1 indicating perfect uniaxial orientation in sheared cylinder-forming copolymers. For example, SAXS analysis of poly(ethylene glycol)-block-poly(ε-caprolactone) micelles reveals core-shell structures with core radii of approximately 10 nm and shell thicknesses of 5 nm, fitted using form factor models to determine aggregation numbers around 50-100.

Applications

Block Copolymer Uses

Block copolymers exhibit microphase-separated morphologies that enable diverse applications by combining the properties of distinct polymer blocks, such as rigidity and elasticity. One prominent use is in thermoplastic elastomers, exemplified by styrene-butadiene-styrene (SBS) triblock copolymers commercialized as Kraton polymers since the 1960s. These materials serve as synthetic rubber alternatives in footwear soles and pressure-sensitive adhesives due to their ability to process like thermoplastics while exhibiting rubber-like elasticity at service temperatures. For instance, SBS-based Kraton D polymers were first produced for footwear in 1964, providing enhanced durability and flexibility compared to vulcanized rubbers. In nanostructured materials, block copolymers act as templates for nanolithography, where polystyrene-block-poly(methyl methacrylate) (PS-b-PMMA) self-assembles into periodic patterns with features as small as 10 nm. This directed self-assembly process enables high-resolution patterning for semiconductor fabrication, surpassing traditional lithography limits by leveraging the block's inherent nanoscale ordering. Studies have demonstrated PS-b-PMMA achieving sub-10 nm half-pitch lines through optimized annealing and etching, critical for next-generation microelectronics. Amphiphilic block copolymers, such as (PEG-b-PLA), form micelles for drug delivery systems, encapsulating hydrophobic therapeutics in their core while the hydrophilic PEG shell enhances solubility and biocompatibility. These micelles enable sustained release of anticancer agents like , improving bioavailability and reducing systemic toxicity in systemic administration. PEG-b-PLA nanoparticles have shown controlled release profiles over days, with encapsulation efficiencies up to 90% for poorly soluble drugs. Recent advances in the 2020s have utilized block copolymer self-assembly to enhance organic photovoltaics, where conjugated blocks form nanostructured active layers that improve charge separation and transport. For example, block copolymers as cathode interlayers in solar cells have improved power conversion efficiencies and stability by optimizing morphology and reducing recombination losses. Performance metrics for block copolymer-based thermoplastic elastomers highlight their mechanical versatility; SBS variants often exhibit tensile strengths of 10-20 MPa and elongations at break exceeding 500%, enabling applications requiring high toughness. In sustainable formulations, such as bio-based thermoplastic elastomers, elongations up to 570% have been achieved alongside tensile strengths around 18 MPa, demonstrating scalability for industrial use.

Engineering and Compatibilization

Copolymers serve as essential compatibilizers in polymer blends by locating at interfaces between immiscible phases, reducing interfacial tension, suppressing coalescence, and stabilizing morphology to achieve finer domain sizes and improved mechanical properties. Block and graft copolymers are particularly effective, with interfacial segments typically comprising 10–15 monomer units for optimal compatibility when solubility parameter differences are ≤0.5 units. For instance, maleic anhydride-grafted (PP-g-MAH) is widely used in (PP-PE) blends, where the polar anhydride groups enhance adhesion to polyethylene phases, leading to reduced domain sizes and enhanced tensile strength. Reactive extrusion enables in-situ compatibilization during melt processing, where copolymers form rapidly (within 5 minutes) through reactions between functional groups in the blend components, improving phase dispersion without pre-synthesis. This technique is common for polyolefin-nylon blends, such as PP-g-MAH reacting with polyamide 6 (PA6) amine end groups to form graft copolymers that stabilize morphology and boost impact resistance. Gradient or graft structures can further refine interfacial properties in such systems. Statistical copolymers are engineered to tune glass transition temperature (Tg) and in materials requiring balanced flexibility and rigidity. copolymers exemplify this, where increasing vinyl acetate content (e.g., >10%) reduces crystallinity, lowers , and enhances flexibility and low-temperature toughness, making EVA suitable for flexible applications like adhesives and films. For impact modification, copolymers like (ABS) are incorporated as tougheners in plastics, leveraging the rubbery phase (5–30 wt%) grafted to a styrene-acrylonitrile backbone to improve and . enhances impact strength (up to 13 ft lb/in) in blends with polycarbonates or terephthalates, enabling durable components in automotive and sectors. In automotive composites, copolymer additives support by enabling lightweight, recyclable structures, such as copolypropylene (CopoPP) blends with recycled and fillers that reduce vehicle weight by 10% and CO2 emissions by 95% compared to virgin materials, while improving by 52%. These post-2000s developments align with eco-friendly goals in interior parts like B-pillars.

Biomedical and Specialty Applications

Copolymers play a pivotal role in biomedical applications due to their tunable properties, such as , biodegradability, and responsiveness to biological environments, enabling advanced therapeutic and diagnostic uses. In , alternating and statistical copolymers form hydrogels that mimic the , providing scaffolds for cell growth and regeneration. For instance, poly(ethylene glycol)-co-acrylamide (PEG-co-acrylamide) hydrogels are widely used as injectable scaffolds in cardiac and , offering high water content, mechanical flexibility, and controlled degradation to support and integration. These hydrogels exhibit thermosensitive or pH-responsive behaviors, allowing gelation and precise delivery of bioactive molecules, with studies demonstrating enhanced cell viability and production in preclinical models. Biodegradable block copolymers, particularly poly(lactic acid)-block-poly(ethylene glycol) (PLA-b-PEG), are essential for resorbable medical devices like sutures and implants, where they provide temporary structural support before hydrolytic degradation into non-toxic byproducts. PLA-b-PEG constructs degrade over weeks to months, depending on the PLA block length and molecular weight, making them suitable for orthopedic fixation devices and drug-eluting implants that reduce the need for secondary surgeries. Their amphiphilic nature also facilitates into nanoparticles for localized drug release, with confirmed through implantation studies showing minimal . In specialty applications, star copolymers serve as efficient non-viral vectors for gene therapy, leveraging their compact, multi-armed architecture to condense DNA or RNA and protect it from enzymatic degradation while enhancing cellular uptake. These cationic star polymers, often based on poly(2-(dimethylamino)ethyl methacrylate), demonstrate superior transfection efficiency compared to linear analogs in vitro and in vivo, with low cytotoxicity due to reduced charge density at the periphery. Similarly, dendritic copolymers enable sensitive biosensors for biomedical diagnostics, where their high surface functionality allows immobilization of recognition elements for detecting biomarkers like glucose or proteins. Linear-dendritic block copolymers, for example, form micellar assemblies that amplify signals in electrochemical or optical sensors, achieving detection limits in the nanomolar range for early disease monitoring. Regulatory frameworks underscore the established safety of certain copolymers in clinical use; poly(lactic-co-glycolic acid) (PLGA), approved by the FDA since the 1970s for sutures and implants, has seen updated assessments in the 2020s confirming its suitability for long-acting systems. Recent evaluations, including a 2020 FDA safety summary, highlight PLGA's low and predictable degradation, supporting its expansion into formulations for sustained release over months. A prominent example of copolymer micelles in cancer therapeutics is Genexol-PM, a PEG-PLA formulation that encapsulates for improved solubility and tumor targeting, approved in and showing response rates of 58-60% in phase II trials for with reduced compared to Cremophor-based alternatives. Clinical data from multicenter studies further validate its efficacy in combination regimens, such as with for non-small cell , demonstrating prolonged . As of 2025, recent advancements include PLGA-based nanoparticles for mRNA delivery in and block copolymer interlayers enabling organic solar cells with efficiencies exceeding 18%.

References

  1. [1]
  2. [2]
    Copolymer - an overview | ScienceDirect Topics
    A copolymer is a polymer formed when two (or more) different types of monomers are linked in the same polymer chain.
  3. [3]
    Advances and applications of block-copolymer-based ...
    In this review, block copolymer classification, synthesis, characterization, stimuli-responsive behavior and nanostructure applications are summarized.
  4. [4]
    copolymer (C01335) - IUPAC Gold Book
    A polymer derived from more than one species of monomer. Note: Copolymers that are obtained by copolymerization of two monomer species are sometimes termed ...
  5. [5]
    [PDF] 7: Source-Based Nomenclature for Copolymers (1985) - iupac
    Mole fractions, or mole percentages, for the monomeric units are placed in parentheses after the copolymer name, followed by the symbol 'x', or the phrase 'mol ...
  6. [6]
    Free Radical Copolymerization
    Composition and sequence length distribution are important in determining the 'blockiness' of the copolymer, microscopic morphology, and macroscopic properties.
  7. [7]
    [PDF] A Brief Guide to Polymer Nomenclature | IUPAC
    Non-linear polymers and copolymers, and polymer assemblies are named using the italicized qualifiers in Table 2. The qualifier, such as branch, is used as a.
  8. [8]
    U.S. Synthetic Rubber Program - National Historic Chemical Landmark
    They discovered in 1929 that Buna S (butadiene and styrene polymerized in an emulsion), when compounded with carbon black, was significantly more durable than ...
  9. [9]
    Styrene-Acrylonitrile Copolymer - an overview | ScienceDirect Topics
    SAN is a solid polymer of 25-30% acrylonitrile and 70-75% styrene, with better resistance to oils and greases than polystyrene.
  10. [10]
    30.1: Chain-Growth Polymers - Chemistry LibreTexts
    Mar 23, 2024 · The mechanism is in most cases a free radical addition. In free radical reactions the pi pair of electrons separates. One of these electrons ...Missing: review | Show results with:review
  11. [11]
  12. [12]
    Introduction to polymers: 4.5.1 The copolymer equation | OpenLearn
    4.3.2 Propagation ... 1 The copolymer equation. It can be shown that the rate of change of monomer concentration in any copolymerization is given by the equation.<|separator|>
  13. [13]
    Section 11.4.1: Ziegler-Natta Polymerizations - Chemistry LibreTexts
    Jun 20, 2023 · The catalyst oxidatively adds an olefin like an ethylene molecule to the coordinatively unsaturated site. This step is followed an olefin ...
  14. [14]
    21.9 Polyamides and Polyesters: Step-Growth Polymers
    Nov 24, 2023 · The polyester Dacron and the polyamide Nylon 66, shown here, are two examples of synthetic condensation polymers, also known as step-growth polymers.
  15. [15]
    Effect of Monomer Solubility on the Evolution of Copolymer ...
    Jan 26, 2017 · In the present work, the effect of monomer solubility on copolymer morphology is explored for an aqueous PISA formulation.
  16. [16]
    Reactivity of monomers in copolymerization - ResearchGate
    Aug 9, 2025 · This selectivity may be attributed to (1) complex formation between monomers or to (2) dipole, steric, or other factors which are independent of ...
  17. [17]
    The Copolymerization of Styrene and Methyl Methacrylate
    Mayo · Frederick M. Lewis. ACS Legacy Archive. Open PDF. Journal of the American Chemical Society. Cite this: J. Am. Chem. Soc. 1944, 66, 9, 1594–1601.
  18. [18]
    [PDF] Chapter 4 Copolymerization
    the cross propagation rate constant for a given radical,. rA = kpAA. kpAB and rB = kpBB. kpBA. (4.41) we can rewrite equation (4.40) as follows: FA. = (rAA + B) ...
  19. [19]
    IUPAC recommended experimental methods and data evaluation ...
    Apr 11, 2024 · ... copolymer compositions can be assumed to be equivalent, allowing direct fitting of the Mayo–Lewis equation (eqn (1)). There is, however, no ...
  20. [20]
    The copolymerization of 1,3‐butadiene with styrene by butyllithium ...
    The copolymer reactivity ratios for initial reaction appear to be 7 for butadiene and nearly 0 for styrene. When the copolymerization is carried out in the ...
  21. [21]
  22. [22]
  23. [23]
    Thermal degradation behaviour of some alternating copolymers
    An overview is presented of the thermal degradation behaviour of a number of alternating copolymers, including maleic anhydride copolymers with vinyl acetate.
  24. [24]
    ZeMac Copolymer Technology Supports Adhesives and Coating ...
    Jun 7, 2018 · ZeMac copolymers are engineered to deliver performance, value and functionality well suited to the adhesive and coating market.
  25. [25]
    periodic copolymer (P04494) - IUPAC
    A copolymer consisting of macromolecules comprising more than two species of monomeric units in regular sequence. Source: PAC, 1996, 68, 2287. (Glossary of ...Missing: definition | Show results with:definition
  26. [26]
    [PDF] Periodic copolymers - Controlled Radical Polymerization
    A good example for the synthesis of periodic copolymers by polymer reactions may be the hydro- genation of alternating butadiene–MMA copolymers by Yokota and ...
  27. [27]
    Preparation of periodic copolymers by living anionic polymerization ...
    Feb 14, 2018 · ... Preparation of periodic copolymers by living anionic polymerization mechanism assisted with a versatile programmed monomer addition mode.
  28. [28]
    Synthetic Rubber - an overview | ScienceDirect Topics
    III.C Product Usages ; SBR, Tires, hoses and belting, footwear, sporting goods, food equipment, textiles, electrical equipment, construction materials, cellular ...
  29. [29]
    Statistical Copolymer - an overview | ScienceDirect Topics
    Statistical copolymers are copolymers in which the sequential distribution of the monomeric units obeys known statistical laws.
  30. [30]
    New Statistical Models for Copolymerization - PMC - NIH
    We introduced two new Markov chain models: the Bernoulli and Geometric models. The major differences to classical copolymer Markov chains based on the ...Missing: statistics | Show results with:statistics
  31. [31]
    Polymers - MSU chemistry
    Addition Copolymerization. Most direct copolymerizations of equimolar mixtures of different monomers give statistical copolymers, or if one monomer is much more ...Missing: Mayo- | Show results with:Mayo-
  32. [32]
    80 years of the Mayo Lewis equation. A comprehensive review on ...
    The combination of r1 >1 and r2 <1, leads to the preferential incorporation of monomer-1 into the copolymer, as addition of monomer-1 is favored over monomer-2 ...Missing: factors | Show results with:factors
  33. [33]
    On estimating reactivity ratios using the integrated mayo–lewis ...
    This work investigates the estimation of reactivity ratios using the integrated Mayo-Lewis equation, particularly in the context of polymerization.<|separator|>
  34. [34]
    (PDF) Interpretations of Polymer-Polymer Miscibility - ResearchGate
    Aug 6, 2025 · For practical purposes, a miscible polymer mixture is defined as a stable homogeneous mixture which exhibits macroscopic properties expected ...
  35. [35]
    [PDF] POLYMERS AND PETROCHEMICALS
    Copolymerization Equation; Mayo-Lewis Equation of Copolymerization o The compositions of statistical copolymers are determined by the monomer composition and.
  36. [36]
    [PDF] STUDY OF ABS AND SAN FAILURES UNDER DROP
    After addition of acrylonitrile, SAN has in comparison with styrene higher impact toughness, strength, fatigue resistance, modulus of elasticity, hardness ...
  37. [37]
    [PDF] Impact Modification of SAN by Dynamic Vulcanization with Liquid ...
    SAN is a random copolymer containing. 60-80% of styrene and 20-40% of acrylonitrile. Among the characteristics of SAN, the important features like rigidity and.Missing: statistical | Show results with:statistical
  38. [38]
    Phase Segregation in Gradient Copolymer Melts - ACS Publications
    Gradient copolymers offer a large degree of control over the A−B interfacial profile, and thus they may be very useful in applications where interfacial ...
  39. [39]
    Gradient Copolymers via In Situ Monomer Transformation with ...
    We developed concurrent tandem living radical polymerization as a novel methodology to efficiently, conveniently, and in one-pot produce gradient copolymers.Gradient Copolymers Via In... · Figure 1 · Scheme 2
  40. [40]
    Gradient copolymers – Preparation, properties and practice
    This review presents an overview of the methods of synthesis of gradient copolymers using the range of available polymerization techniques, and compares and ...
  41. [41]
    Synthesis and characterization of functional gradient copolymers of ...
    Epoxy-functional spontaneous gradient copolymers of glycidyl methacrylate (G) and n-butyl acrylate (B) were synthesized via atom transfer radical ...<|separator|>
  42. [42]
  43. [43]
    Tacticity Changes during Controlled Degradation of Polypropylene
    Sep 21, 2021 · The controlled reactive degradation of polypropylene (CPP) in a twin-screw extruder has been investigated from the perspective of tacticity.
  44. [44]
    Oscillating Stereocontrol: A Strategy for the Synthesis of ... - Science
    A strategy has been developed for the synthesis of thermoplastic elastomeric polypropylene based on the catalytic activity of the unbridged metallocene ...<|control11|><|separator|>
  45. [45]
    Synthesis of stereoblock polypropylene by change of temperature in ...
    Synthesis of stereoblock polypropylene by change of temperature in living polymerization. Articles; Published: 14 August 2010. Volume 18, pages 737–741, ( ...
  46. [46]
    Synthesis and Characterization of Stereoblock Poly(lactic acid)s with ...
    PLLA is a melt-processable, crystalline polymer, and its fibers and films show mechanical properties comparable to those of polystyrene and polypropylene. 1,2 ...
  47. [47]
    Graft copolymers - Matyjaszewski Polymer Group
    Graft copolymers belong to the general class of segmented copolymers and generally consist of a linear backbone of one composition and randomly distributed ...
  48. [48]
    Well-defined graft copolymers: from controlled synthesis to ...
    Nov 25, 2010 · In this tutorial review, recent advances in synthetic approaches towards the construction of well-defined graft copolymers are discussed in detail
  49. [49]
    Polymer Grafting and its chemical reactions - Frontiers
    There are three general methods for synthesizing graft copolymers: 1) “grafting to,” 2) “grafting from,” and 3) “grafting through” it is also summarized in ...
  50. [50]
    Poly(ε-caprolactone)-Based Graft Copolymers: Synthesis Methods ...
    Grafting of polymer chains is mainly carried out by “grafting from” or “grafting onto” methods. ... Synthesis of a PCL-g-PLA according to a “grafting through” ...
  51. [51]
    PMMA-g-OEtOx Graft Copolymers: Influence of Grafting Degree and ...
    Mar 30, 2018 · Depending on the degree of grafting (DG) and the side chain degree of polymerization (DP), graft copolymers may feature properties similar ...
  52. [52]
    A Review on the Synthesis, Characterization, and Modeling ... - MDPI
    ... grafting to” and “grafting through” polymer grafting techniques, respectively. ... Grafting, from; grafting to; network copolymer grafting, Grafting from ...
  53. [53]
    [PDF] 1 Toughening Polylactide with Graft-Block Polymers Bongjoon Lee ...
    Mixing PLA with graft-block polymers (g-ML) enhances its tensile toughness, achieving 280% strain at break compared to 7% for pure PLA.
  54. [54]
    Compatibility Effect of Added Graft Copolymer on Polystyrene and ...
    It is clear that the graft copolymer added here has more marked effect not only on impact strength but on tensile and bending strengthes as well. Figure I shows ...
  55. [55]
    Unraveling the Particle Morphology of ABS Polymer Latexes by 3D ...
    Aug 6, 2025 · The grafted ABS latex was synthesized by seeded semibatch emulsion polymerization using a water-soluble initiator. The seed was a PB latex, and ...
  56. [56]
    Star Polymers | Chemical Reviews
    The most attractive aspect of the core-first approach is its excellent yields, as pure star polymers can be conveniently isolated from the crude reaction ...
  57. [57]
    Star Polymer - an overview | ScienceDirect Topics
    Star polymers are branched polymers with multiple arms or branches radiating from a central core, consisting of linear chains linked to a central core.
  58. [58]
    Synthesis of polyethylene and polystyrene miktoarm star copolymers ...
    Miktoarm star copolymers (MSPs) having multiple polyethylene (PE) and polystyrene (PS) arms joined at the crosslinked polydivinylbenzene (PDVB) core were ...
  59. [59]
    Application of star poly(ethylene glycol) derivatives in drug delivery ...
    Jul 10, 2020 · Multi-arm star poly(ethylene glycol) (PEG) polymers have a number of advantages over linear PEG polymers when used as carriers for drug delivery and controlled ...Missing: oxide) | Show results with:oxide)
  60. [60]
    Dendrimer-based bionanomaterials produced by surface ... - Nature
    Apr 18, 2012 · We compared thermosensitive properties of PAMAM dendrimers having N-isopropylamide (NIPAM) groups on all chain terminals with those of poly( ...Peg-Modified Pamam... · Dendrimer-Based Amphiphiles... · Dendrimer--Gold Np Hybrids
  61. [61]
    Precise Synthesis of Dendrimer‐Like Star‐Branched Polymers, a ...
    Apr 27, 2011 · These polymers are very similar in branched architecture to dendrimers, but are composed of many polymer chains between the junctions in all ...
  62. [62]
    Dendrimer-based drug delivery systems: history, challenges, and ...
    Jul 25, 2022 · This mini review article provides a brief overview of dendrimer's history and properties and the latest developments of dendrimers as drug delivery systems.
  63. [63]
    Impact of Dendrimers on Solubility of Hydrophobic Drug Molecules
    The effect of dendrimer generation at constant pH increased the solubility in an order of G3 > G2 > G1 > G0. Effect of dendrimer pH on the solubility ...
  64. [64]
    Dendrimers: Exploring Their Wide Structural Variety and Applications
    Dendrimers are hyper-branched macromolecules characterized by large numbers of end-group functionalities and a compact molecular structure.
  65. [65]
    Synthesis of Nonionic Dendrimer-like Star Block Copolymers Based ...
    Aug 6, 2025 · Nabid et al. synthesized a nonionic dendrimer-like star block copolymer of PCL and PEG, which was used as a stabilizer for the preparation of ...
  66. [66]
    Synthesis and Characterization of Star Poly(ε-caprolactone)-b ... - NIH
    Most of the work has focused on classical micelles formed by intermolecular aggregation of amphiphilic block copolymers as the drug delivery vehicle, and the ...
  67. [67]
    Theory of Microphase Separation in Block Copolymers
    Galactose-Based Block Copolymers with Switchable χ Parameters for Tuning Phase Separated Morphology. ... Flory–Huggins Interaction Parameter for Next ...
  68. [68]
    Crazing in polystyrene-polybutadiene diblock copolymers ...
    Polystyrene-polybutadiene diblock copolymers exhibiting morphologies of cylindrical rubbery domains in a glassy matrix were studied in tensile stress-strain ...
  69. [69]
    Macrophase versus Microphase Separation in Solutions of Block ...
    Sep 2, 2020 · The aim of this work is an in-depth study of the stability of solutions of block copolymers with respect to both micro- and macrophase ...
  70. [70]
    Topology Effect on Order–Disorder Transition of High-χ Block ...
    Jul 18, 2024 · This work aims to systematically examine the topology effect on the self-assembly of block copolymers.
  71. [71]
    [PDF] Shear-Aligned Block Copolymer Monolayers as Seeds To Control ...
    Sep 27, 2016 · As annealing proceeds, the interaction between the shear- aligned and nonaligned layers drives the complete annihilation of disclinations and ...
  72. [72]
    Polymer Blend Compatibilization by Gradient Copolymer Addition ...
    Polymer Blend Compatibilization by Gradient Copolymer Addition during Melt Processing: Stabilization of Dispersed Phase to Static Coarsening | Macromolecules.Missing: widen | Show results with:widen
  73. [73]
    Carbon-13 Nuclear Magnetic Resonance Determination of ...
    Carbon-13 Nuclear Magnetic Resonance Determination of Monomer Composition and Sequence Distribution in Ethylene-Propylene Copolymers Prepared with a ...
  74. [74]
    [PDF] 77Se NMR as a Tool to Examine Copolymer Seque
    Jul 29, 2022 · Finally, the NMR spectra enable estimation of reactivity ratios for the copolymerization. Copolymer reactivity ratios are often determined by ...
  75. [75]
    Broadening of carbonyl stretching vibration bands appearing for ...
    Bands Appearing for Acrylate Copolymers. Sir: Environmental effects upon infrared and nmr spectra have been discussed for the purpose of in- vestigating the ...
  76. [76]
    Feasibility of the Simultaneous Determination of Monomer ...
    Dec 7, 2015 · An immersion Raman probe was used in emulsion copolymerization reactions to measure monomer concentrations and particle sizes.
  77. [77]
    Infrared, Raman, and Near-Infrared Spectroscopic Evidence for the ...
    Fourier-transform infrared (FT-IR), near-infrared (NIR)-excited FT-Raman, and FT-NIR spectra have been measured for poly(acrylic acid) (PAA) in a cast film.
  78. [78]
    Conjugated alternating copolymers containing both donor and ...
    A straightforward synthesis toward two conjugated alternating copolymers consisting of 2,7-linked carbazole donor (2,7-Cz) and ladderized pentaphenylene with ...
  79. [79]
    Synthesis of Light-Emitting Conjugated Polymers for Applications in ...
    (220) The UV−vis spectrum of copolymer 5 prepared in this way indicates that the polymer contains conjugated sequences mainly comprising stilbene and ...
  80. [80]
    Polymer Molecular Weight Analysis by 1H NMR Spectroscopy
    Demonstrated herein are the simplicity, reproducibility, and convenience of 1H NMR spectroscopy in the analysis of polymer number-average molecular weight (Mn), ...
  81. [81]
    End Group Analysis Accounts for the Low Molecular Weight ...
    End Group Analysis Accounts for the Low Molecular Weight Observed in the 1,3-Divinyltetramethyldisiloxane−Pt Complex Catalyzed Hydrosilylation Copolymerization ...
  82. [82]
    Carbon-13 Nuclear Magnetic Resonance Characterization of ...
    Polymerization of coordinated monomers. XVIII. Sequence distribution in methyl acrylate–styrene copolymers prepared in the presence of zinc chloride.
  83. [83]
    Priming self-assembly pathways by stacking block copolymers
    Nov 14, 2022 · We developed a pathway priming strategy combining control of thin film initial configurations and ordering history.
  84. [84]
    Styrenic Block Copolymers (SBCs) - Kraton Corporation
    Styrene Blocks: These blocks are rigid and contribute properties such as strength, thermal stability, and resistance to chemicals.
  85. [85]
    Styrenic block copolymer‐based thermoplastic elastomers in smart ...
    Jul 29, 2022 · The cyclic block copolymers offer much lower hydrodynamic volume, phase-separated structure with block sizes less than 20 nm, lower viscosities ...Abstract · INTRODUCTION AND... · ADVANCEMENTS IN... · IMPACTS OF...<|control11|><|separator|>
  86. [86]
    Kraton Polymers LLC Employer Profile - NSBE Career Center
    In 1964, Kraton's Torrance facility made the first production of Kraton D polymers and SBS compounds for footwear applications. About the same time the ...
  87. [87]
    Chemically tailored block copolymers for highly reliable sub-10-nm ...
    Jul 6, 2024 · We report the creation of a chemically tailored, highly reliable, and practically applicable block copolymer and sub-10-nm line patterns by directed self- ...Results · Ps-B-Pgm Synthesis And... · Bulk Morphologies
  88. [88]
    Sub-10 nm Feature Sizes of Disordered Polystyrene-block-poly ...
    Dec 23, 2019 · Weak segregation of polystyrene-block-poly(methyl methacrylate) block copolymers (PS-b-PMMA BCPs) limits their utility for sub-10 nm lithography
  89. [89]
    Scaling-down lithographic dimensions with block-copolymer materials
    Oct 25, 2025 · We demonstrate in particular that this kind of polymer can efficiently achieve periodic features close to 10 nm. These thin-films can be ...
  90. [90]
    PEG-b-PLA Micelles and PLGA-b-PEG-b-PLGA Sol-Gels for Drug ...
    PEG-b-PLA micelles are a first-generation platform for the systemic multi-delivery of poorly water soluble anticancer agents. PLGA-b-PEG-b-PLGA sol-gels are a ...2. Peg-B-Pla Micelles · 2.2. Drug Release From... · 3. Plga-B-Peg-B-Plga...
  91. [91]
    Recent advances in PEG–PLA block copolymer nanoparticles
    Nov 26, 2010 · In many studies, PEG–PLA block copolymer nanoparticles were used as carriers for vaccine, protein, and gene drugs, particularly in a sustained/ ...
  92. [92]
    Different Insight into Amphiphilic PEG-PLA Copolymers: Influence of ...
    Amphiphilic block poly(propylene carbonate)‐block‐allyloxypolyethyleneglycol copolymer based shell cross‐linked micelles for controlled release of drug.
  93. [93]
    Block copolymers as efficient cathode interlayer materials for ...
    Jan 25, 2021 · Block copolymers have high optical transmittance, low work function, excellent conductivity, and provide higher power conversion efficiency and ...<|separator|>
  94. [94]
    Highly Impact-Resistant Block Polymer-Based Thermoplastic ...
    Dec 19, 2021 · Tensile stress at a strain of 300%. e. Tensile strength. f. Elongation at break. g. Tensile toughness estimated from the inner area of a stress ...
  95. [95]
    [PDF] Sustainable thermoplastic elastomers produced via cationic RAFT ...
    Jan 13, 2021 · Specifically, PDHF-0.23 exhibited the highest elongation observed in a PVE TPE (570 ± 70%). This study demonstrates a significant advance in the ...
  96. [96]
    Fully recyclable and tough thermoplastic elastomers from simple bio ...
    Sep 10, 2024 · Uniaxial tensile testing reveals that TPE-8 has good tensile strength, elongation at break and cyclic tensile results (σB = 17.8 ± 0.9 MPa ...<|control11|><|separator|>
  97. [97]
    Compatibilization phenomenon in polymer science and technology
    Compatibilization by polymers formed in situ by reaction between active and co-reactive groups in blending polymers is a commonly used strategy for ...
  98. [98]
    Ethylene-Vinyl Acetate - an overview | ScienceDirect Topics
    Ethylene-vinyl acetate (EVA) is defined as a polar copolymer of ethylene and vinyl acetate that offers increased flexibility, elongation, ...
  99. [99]
    Acrylonitrile Butadiene Styrene - an overview | ScienceDirect Topics
    ABS has excellent impact strength that is higher than all other plastics besides PC. Polycarbonate (PC) and ABS can be blended to form an excellent polymer ...Processing Recommendations... · 22.1 Acrylonitrile Butadiene... · Material Profiles
  100. [100]
    Sustainable Engineered Designs and Manufacturing of Waste ...
    Aug 3, 2024 · This study presents a sustainable and effective approach for lightweight automotive interior part production using a synergistically designed PP compound.
  101. [101]
    Dendrimers: synthesis, applications, and properties - PMC
    Dendrimers are nano-sized, radially symmetric molecules with well-defined, homogeneous, and monodisperse structure consisting of tree-like arms or branches [1].
  102. [102]
    Injectable Hydrogels for Cardiac Tissue Engineering - PMC - NIH
    ... polymers to PEG hydrogels or incorporated bioactive molecules. For example, Chow et al. used PEG acrylate copolymerized with PEG dithiol to develop a ...
  103. [103]
    Responsive Acrylamide-Based Hydrogels: Advances in ... - NIH
    Jun 21, 2024 · Hydrogels, composed of hydrophilic homopolymer or copolymer networks, have structures similar to natural living tissues, making them ideal ...
  104. [104]
    Biodegradable Medical Implants: Reshaping Future Medical Practice
    Aug 21, 2025 · Synthetic polymers such as PLA, PGA) and their copolymers (PLGA, PCL) deliver tunable strength and degradation rates, while natural ...
  105. [105]
    Review Medical applications and prospects of polylactic acid materials
    Dec 20, 2024 · PLA polymers have been utilized in orthopedics for the fabrication of biodegradable screws, fixation pins, plates, and suture anchors, which ...
  106. [106]
    Biodegradable PEG-Based Amphiphilic Block Copolymers for ...
    This review discusses the design and applications of amphiphilic block copolymer scaffolds integrating hydrophilic poly(ethylene glycol) (PEG) blocks with ...
  107. [107]
    Star polymers for gene delivery - Georgiou - Wiley Online Library
    Feb 3, 2014 · Star polymers offer great advantages in gene delivery compared to linear and branched polymers. These are summarized and discussed in this ...Introduction · Synthetic Methodologies · Structure--Activity...
  108. [108]
    Micelle Forming Linear–Dendritic Block Copolymers: A Theoretical ...
    Jun 1, 2025 · This makes them ideal candidates for engineering nanoparticles with tailored properties for applications such as drug delivery and sensing. Here ...
  109. [109]
    PLGA-Based Nanomedicine: History of Advancement and ... - NIH
    PLGA are a family of FDA-approved biodegradable polymers that are physically strong and highly biocompatible and have been extensively studied as delivery ...Missing: 2020s | Show results with:2020s
  110. [110]
    [PDF] Medical Device Material Safety Summaries - ECRI Reports - FDA
    Nov 9, 2020 · The Safety Brief summarizes the findings of the literature search on toxicity/biocompatibility of P(L/G)A. ... Results for PLGA/ACP copolymers ...Missing: history 2020s
  111. [111]
    Polymeric Micelles in Anticancer Therapy: Targeting, Imaging ... - NIH
    In phase II studies, Genexol-PM was found to be effective and safe with high response rates in patients suffering from metastatic breast cancer and advanced ...
  112. [112]
    Multicenter Phase II Clinical Trial of Genexol-PM® with Gemcitabine ...
    This multicenter phase II clinical trial was designed to evaluate the efficacy and safety of weekly Genexol-PM® and gemcitabine combination chemotherapy